113
Copyright by Luca Magenes 2011

Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

Copyright

by

Luca Magenes

2011

Page 2: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

The Thesis Committee for Luca Magenes

Certifies that this is the approved version of the following thesis:

Fatigue Assessment of High Mast Illumination Poles

Using Field Measurements

APPROVED BY

SUPERVISING COMMITTEE:

Todd Helwig

Karl Frank

Supervisor:

Page 3: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

Fatigue Assessment of High Mast Illumination Poles

Using Field Measurements

by

Luca Magenes B.S.

Thesis

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science in Engineering

The University of Texas at Austin

August 2011

Page 4: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

Dedication

I dedicate my work to my family and to Valeria for the continuous love and support

during these two years.

Page 5: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

v

Acknowledgements

I would like to acknowledge Dr. Helwig and Dr. Frank for the support and

motivation, my friends and colleagues Jeremiah Fasl and James Kleineck for the

substantial contribution to this work, the sponsor of the project Texas Department of

Transportation, and personally Michael Smith and Steve Eveland for the support during

the field instrumentation trips. I also would like to thank Vasileios Samaras and Matt

Reichenbach for taking part in one of the field trips and each one that helped me out in

every little thing during the past two years at Ferguson Lab.

August 2011

Page 6: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

vi

Abstract

Fatigue Assessment of High Mast Illumination Poles

Using Field Measurements

Luca Magenes, M.S.E.

The University of Texas at Austin, 2011

Supervisor: Todd A. Helwig

Failures of high mast illumination poles (HMIPs) in recent years have raised

concerns on the long-term fatigue performance of the poles by various transportation

officials around the US. The thesis documents a study sponsored by the Texas

Department of Transportation focused on the fatigue behavior of in-service HMIP

systems. This study is an extension of previous investigations on the fatigue behavior of

the poles that have demonstrated that many poles have poor performance and fail in

fatigue before the AASHTO category E' limit. Galvanized specimens were also tested

and some of them showed evidence of initial cracking, impacting the fatigue performance

such that the galvanized poles behaved worse than the uncoated specimens.

Ultrasonic Testing (UT) has shown several poles around the state of Texas contain cracks

in the welds between the shaft and base plate. To further investigate the performance of

the poles in-service, a field study was initiated to measure the wind speed and direction,

as well as the corresponding stresses in the pole shaft. This thesis presents results from

the field investigation.

Page 7: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

vii

A data acquisition system was developed to gather wind data and induced stresses. The

system was powered by a solar panel and can be remotely accessed via a wireless

modem. Data collected throughout the year details the intensity and number of stress

cycles experienced by the poles, and could be correlated with the measured wind

velocity. Using the field data, more accurate estimates of expected fatigue life for the

poles were made. The study provides TxDOT with valuable data on the performance of

in-service poles so that the most critical fatigue cases can be identified and proper

decisions can be made on the appropriate inspection or repair schedule.

Page 8: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

viii

Table of Contents

Chapter 1 Introduction......................................................................................................1

1.1 Description of the problem ......................................................................................1

1.2 High Mast Illumination Poles Failures ....................................................................2

1.3 Scope of the Study and this Thesis ..........................................................................5

Chapter 2 Previous Research ............................................................................................7

2.1 Load Testing of High Mast Illumination Poles........................................................7

2.1.1 Phase I (Rios) ..................................................................................................8

2.1.2 Phase II (Stam) ..............................................................................................10

2.1.3 Phase III (Pool) .............................................................................................13

2.2 Thermal Investigation during Galvanization .........................................................16

2.3 Wind Effects on High Mast Illumination Poles .....................................................17

2.3.1 Vortex Shedding ...........................................................................................18

2.3.2 Buffeting .......................................................................................................19

2.3.3 Static Parameters ...........................................................................................19

2.3.4 AASHTO Wind Design Specifications ........................................................20

2.3.5 Iowa Study ....................................................................................................22

Chapter 3 Field Instrumentation ....................................................................................28

3.1 Objective of the Instrumentation ...........................................................................28

3.2 Hardware ................................................................................................................31

3.2.1 Stress Measurement ......................................................................................33

3.2.2 Wind Measurement .......................................................................................33

3.2.3 Data Acquisition and Communication ..........................................................35

3.2.4 Powering the System.....................................................................................35

3.3 Software .................................................................................................................36

3.3.1 Rainflow ........................................................................................................36

Page 9: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

ix

3.3.2 Event Capture................................................................................................37

3.3.3 Voltage Measurement ...................................................................................37

3.3.4 Remote Access and Control ..........................................................................38

Chapter 4 Background Information for Field Studies .................................................40

4.1 Introduction ............................................................................................................40

4.2 TxDOT Ultrasonic Inspections ..............................................................................40

4.2.1 UT Results on Texas In-Service HMIPs ......................................................41

4.2.2 Cracking on Instrumented HMIPs ...............................................................42

4.2.3 Cracking on Instrumented HMIPs ...............................................................45

4.3 Wind Historical Records ........................................................................................46

4.4 Summary ................................................................................................................52

Chapter 5 Data Analysis ..................................................................................................53

5.1 Supporting Finite Element Modeling.....................................................................53

5.1.1 Beam Model ..................................................................................................53

5.1.2 Shell Model ...................................................................................................58

5.2 Event Capture Analysis .........................................................................................61

5.2.1 Data Observation ..........................................................................................61

5.2.2 Frequency Analysis .......................................................................................66

5.2.3 Wind Speed to Stress Correlation .................................................................70

5.3 Fatigue Assessment with Rainflow Data ..............................................................74

5.3.1 Cumulative Damage Theory ........................................................................75

5.3.2 Definition of fatigue performance ................................................................76

5.3.3 Estimates of the fatigue life .........................................................................78

5.4 Modification to the estimate due to short term measurement ...............................85

Chapter 6 Conclusions .....................................................................................................90

6.1 Introduction ...........................................................................................................90

Page 10: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

x

6.2 Finite Element Analyses ........................................................................................90

6.3 Field Monitoring Data............................................................................................90

6.4 Future Work ...........................................................................................................91

6.5 Recommendations to HMIP Owners .....................................................................91

Appendix A ......................................................................................................................93

References .........................................................................................................................96

Vita ....................................................................................................................................98

Page 11: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

xi

List of Tables

Table 2.1: Phase III HMIP fatigue testing specimens (Pool ,2010) ...................14

Table 2.2: Ultrasonic Tests Results after Galvanization ....................................17

Table 2.3: Modal Frequencies and Damping Ratios fir the Iowa HMIP ...........24

Table 4.1: TxDOT Ultrasonic Testing Results ..................................................41

Table 4.2: UT results for the field-instrumented HMIPs ...................................44

Table 4.3: UT results on HMIP specimens before load testing .........................45

Table 4.4: Average Wind speeds in the Instrumented HMIP locations .............47

Table 4.5: Maximum gusts speeds in the HMIP locations.................................47

Table 5.1: Vortex Shedding Critical Wind Speed for the First Three Natural Modes

...........................................................................................................55

Table 5.2: Influence of anchor bolts on fatigue life ...........................................60

Table 5.3: Comparison between the FEM frequencies and FFT frequencies ....67

Table 5.4: Texas HMIP natural frequencies vs. Iowa HMIP natural frequencies 68

Table 5.5: AASHTO A values for the fatigue performance ...............................76

Table 5.6: Fatigue life estimates for the El Paso US-54 HMIP .........................82

Table 5.7: Fatigue life estimates for the El Paso IH-10 HMIP ..........................82

Table 5.8: Fatigue life estimates for the Lubbock HMIP...................................83

Table 5.9: Fatigue life estimates for the Corpus Christi HMIP .........................83

Table 5.10: Modification of the fatigue life estimate for short term monitoring period

...........................................................................................................88

Table 5.11: Modification of the fatigue life estimate for short term monitoring period 89

Page 12: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

xii

List of Figures

Figure 1.1: High Mast Illumination Poles in Houston, TX ...................................1

Figure 1.2: Section view of the shaft to base plate connection .............................2

Figure 1.3: Microscope picture of the crack location ............................................3

Figure 1.4: HMIP collapse in Sioux City, Iowa ....................................................4

Figure 1.5: Baseplate of the Collapsed HIMP in Sioux City, Iowa ......................4

Figure 1.6: HIMP Collapse in Rapid City, South Dakota .....................................4

Figure 1.7: Collapse of a stadium sighting Pole over a gymnasium .....................5

Figure 1.8: Cracks above the weld toe in stadium sighting poles .........................5

Figure 2.1: Socket connection (top left), Texas connection (top right), Wyoming

connection (bottom) ............................................................................9

Figure 2.2: Load Testing Setup and Moment diagram ........................................10

Figure 2.3: Texas detail with External Collar .....................................................11

Figure 2.4: Wyoming detail with External Collar ...............................................11

Figure 2.5: Extrapolation of Hotspot Stress from Finite Element Model ...........12

Figure 2.6: Fatigue Test Results ..........................................................................15

Figure 2.7: HMIP specimen entering the zinc bath .............................................16

Figure 2.8: FEM for Thermal Sudies ..................................................................16

Figure 2.9: Drag Coefficient for a Dodecagonal section as suggested by AASHTO,

2006...................................................................................................21

Figure 2.10: Wind Tunnel Testing Setup from the Iowa Study ............................23

Figure 2.11: Measured Drag Coefficient for a 12-sided section ...........................23

Figure 2.12: Measured Stress Range vs. Wind Speed from Iowa Study ...............25

Page 13: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

xiii

Figure 2.13: Probability Distributions of Mean Wind Speed in Sioux City .........26

Figure 2.14: Predicted response with a buffeting and vortex shedding integrated

mathematical model ..........................................................................27

Figure 3.1: Location of instrumented HMIPs on a wind activity map ................29

Figure 3.2: Typical 12 Sided Pole instrumented (Austin, TX)............................30

Figure 3.3: Scheme of the Instrumentation: ........................................................31

Figure 3.3: Plan view of the instrumentation on IH-10 (El Paso, TX): ...............32

Figure 3.5: Wind Monitor Model Young 05103, Helicoid Propeller Anemometer

...........................................................................................................34

Figure 3.6: Setting up the anemometer on a pole (El Paso, TX) .........................35

Figure 3.7: Solar panel and wireless antenna mounted on a pole (El Paso, TX) 35

Figure 3.8: Stress History and Rainflow Counting Example ..............................37

Figure 3.9: Screenshot of the Instrumentation Program ......................................39

Figure 4.1: TxDOT Ultrasonic Testing Results ..................................................42

Figure 4.2: Schematic representation of differential heating due to sun radiation

heating of the HMIP..........................................................................44

Figure 4.3: Rate of occurrence of the daily average wind speed at the HMIP

Locations ...........................................................................................48

Figure 4.4: Yearly wind rose for Corpus Christi (Texas Commission on

Environmental Quality) ....................................................................49

Figure 4.5: Yearly wind rose for El Paso (Texas Commission on Environmental

Quality) .............................................................................................50

Page 14: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

xiv

Figure 4.6: Yearly wind rose for Lubbock (Texas Commission on Environmental

Quality) .............................................................................................51

Figure 5.1: Natural Periods of the HMIP from the Finite Element Model..........54

Figure 5.2: Bending moment shapes for the natural modes of the HMIP ...........54

Figure 5.3: Displacement shape (left) and Moment diagram (right) due to sun

temperature gradient and P-delta effects...........................................57

Figure 5.4: Shell Finite Element Model ..............................................................59

Figure 5.5: Shell stresses S22 [ksi]......................................................................60

Figure 5.6: Event recorded on the El Paso US-54 pole at a 50mph wind speed 62

Figure 5.7: Event recorded on the El Paso US-54 pole at a 7 mph wind speed

direction S-S-W ................................................................................63

Figure 5.8: Event recorded on the Lubbock pole at a 25 mph wind speed, direction

W-N-W .............................................................................................64

Figure 5.9: Event recorded on the Lubbock pole at a 10 mph wind speed, direction

S-W ...................................................................................................65

Figure 5.10: Fourier Spectrum of a strain gage record ..........................................67

Figure 5.11: Fourier Spectrum (FFT) and cumulative RMS of a strain gage record 68

Figure 5.12: Cumulative RMS of stress records at different triggering wind speeds

...........................................................................................................69

Figure 5.13: Max measured Stress vs. 3 second wind speed .................................71

Figure 5.14: Simplified model for the equivalent wind push ................................72

Figure 5.15: Max measured Stress vs. 1 minute wind speed .................................74

Figure 5.16: Definition of the Fatigue Performance of a cracked HMIP based on load

test results..........................................................................................78

Page 15: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

xv

Figure 5.17: Rainflow Counting per day of record ...............................................79

Figure 5.18: Fatigue damage contribution of each stress range ............................81

Figure 5.19: Fatigue Life Estimate (cracked) as function of the Constant Amplitude

Limit Stress for the instrumented HMIPs .........................................84

Figure 5.20: Days of Operation and Average Monthly wind speed for the Corpus Christi

HMIP .................................................................................................86

Figure 5.21: Days of Operation and Average Monthly wind speed for the El Paso

Lubbock ............................................................................................86

Figure 5.22: Days of Operation and Average Monthly wind speed for the El Paso US-

54 HMIP............................................................................................87

Figure 5.23: Days of Operation and Average Monthly wind speed for the El Paso IH-

10 HMIP............................................................................................87

Page 16: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

1

Chapter 1 Introduction

1.1 Description of the problem

High Mast Illumination Poles (HMIPs) as shown in Figure 1,1. are multisided, hollow, cantilever

poles that are typically used to provide light to highway intersections and other urban

environments. The poles are typically 70 to 175 ft (22m to 54m) tall that are anchored to a

foundation and support a lighting frame on top. The details and specifications for the design of

these poles typically vary among different states.

Fig. 1.1 High Mast Illumination Poles in Houston, TX (courtesy of http://www.mckeehen.net/)

Page 17: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

2

Following collapses of HMIPs in the US, The Texas Department of Transportation (TxDOT)

began a campaign of ultrasonic inspections of in-service poles. TxDOT currently has pole designs

with either 8-sided or 12-side pole shafts, depending on the pole height and design wind speed.

The connection details between the shaft and the base plate can either be a full penetration weld

or might also use a ground sleeve, which is a doubler plate that goes around the shaft as shown in

Figure 1.2 (left). Though no collapses have occurred in Texas, several poles have been found with

cracks at the shaft to base plate connection. The twelve-sided, 150 ft (46 m) high pole, designed

for 80 mph (36 m/s) wind speed without a ground sleeve performed very poorly in laboratory

tests (Pool, 2010). In these poles, the shaft is connected to the base plate with a full penetration

weld that has a 3/8" (9.5 mm) root opening as shown in Figure 1.2 (right). Cracks were most

prevalent near the bends of the shaft, at the top of the weld toe as shown in the photograph taken

through the microscope in Figure 1.3. TxDoT has found that 100% of the inspected poles with

these design features possessed cracks at one or more bends.

Fig. 1.2. Section view of the shaft to base plate connection, with the external collar (left) and

without the external collar (right)

Page 18: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

3

Fig. 1.3 Microscope picture of the crack location

1.2 High Mast Illumination Poles Failures

There have been a number of collapses of HMIPs reported around the US. A 140 ft tall HMIP

along Interstate 29 in Sioux City, Iowa fractured at the shaft to baseplate joint in November of

2003 as shown in Figs. 1.4 and 1.5. The pole had a 1 ¼" thick baseplate, and a 3/16" thick shaft

with a twelve-sided section.

Page 19: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

4

Fig. 1.4. HMIP collapse in Sioux City, Iowa

(MTC, 2007)

Fig. 1.5. Baseplate of the Collapsed HIMP in

Sioux City, Iowa (Warpinski, 2007)

In November 2005 the 150 ft tall HMIP shown in Fig. 1.6 collapsed in Rapid City, South Dakota.

The pole had a 1 ¾" thick baseplate and a 3/8” thick shaft. The pole had been installed only two

years prior to the collapse.

Fig.1.6. HIMP Collapse in Rapid City, South Dakota (Rapid City Journal)

Fractures have also been reported on in illumination poles for recreational parks, schools, and

outdoor stadiums. Many of these poles have heights ranging from 70 ft. to 135 ft. Nine incidents

of these Stadium lighting poles have been confirmed by the Costumer Product Safety

Commission from 2000 to 2006. One involved a pole falling through a roof of a school

Page 20: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

5

gymnasium (Fig. 1.7), causing significant property damage. CPSC also confirmed that 50 other

poles with this design developed fracture or cracks next to the weld joint with the base plate

similar to the crack shown in Fig. 1.8.

Fig.1.7. Collapse of a stadium sighting Pole over a gymnasium (Left), and in a practice field

(Right) (CPSC website)

Fig.1.8. Cracks above the weld toe in stadium sighting poles (CPSC website)

A common factor in these collapses is the crack location on the shaft directly above the weld toe.

Previous investigations have confirmed the fatigue behavior of such cracks (Rios, Stam, Pool).

1.3 Scope of the study

The research documented in this thesis is part of a larger study focusing on both the cracking

behavior during the galvanizing process as well as the in-service behavior of poles monitored

Page 21: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

6

through field instrumentation. Results from the study during the galvanizing process are

presented by Kleineck (2011), while this thesis focuses on the field monitoring.

The focus of the field instrumentation is to measure the response of HMIP’s subjected to wind

events at various locations throughout the state of Texas. The data targeted in the instrumentation

included the speed and direction of the wind along with the corresponding stresses in the pole

shaft. The wind and strain data can be compared to laboratory test results to provide a better

estimate of the remaining fatigue life. Several High Mast Illumination Poles were instrumented

around the State of Texas. The poles designs are similar, featuring a total height of 150 ft and a

base diameter of 33. The 3/8 thick pole shafts were 12-sided and had a 33 thick based

diameter welded to a 3 thick baseplate. One of the goals of the instrumentation was to monitor

the behavior at various times through the year to obtain a better measure of the long-term

performance. This long term data provides an indication of the amount of loading that the

structures are subjected to during their service lives so that a fatigue life estimate can be carried

out for the poles. The data is correlated to load test results obtained from previous work

conducted at Ferguson Structural Engineering Laboratory at The University of Texas at Austin.

Wind direction and velocity were also recorded so that a correlation between stress on the

structure and the aerodynamic excitation could be made.

The data gathered over time was used to define a stress history that can be used to estimate an

equivalent average stress that the HMIP will be subjected to during its service life. The use of this

data with load test results are used to provide a good estimate of the remaining fatigue life for the

poles of various designs and differing support conditions. These estimates will help TxDOT

identify the most critical HMIP’s that might require replacement, repair, or more frequent

ultrasonic inspections. The experience gathered from this long-term instrumentation will also

contribute to the development of reliable, remote data acquisition systems for long-term structural

monitoring.

Page 22: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

7

Chapter 2 Previous Research

2.1 Introduction

There have been a number of previous investigations on the behavior of HMIP systems. Because

there have been a number of problems reported on the fatigue behavior of HMIP baseplate

connections, much of the recent research has focused on the fatigue performance of various

connection details. Previous research pointed out how the HIMP failures occurring around the

US were fatigue type failures (Rios, 2007), and that the cracking initiated at the shaft to baseplate

connection, at the bends of the shaft (Rios, 2007). This chapter provides a summary of the recent

research as well as other background information necessary for understanding the material

presented later in this thesis. The research discussed in this thesis is part of a series of research

investigations that have been conducted at the University of Texas over the past several years.

The following subsections focus on the results of these past studies. An overview of the dynamic

behavior of HMIP sections along with background information on the impact of wind loads on

pole sections is then provided.

2.2 University of Texas HMIP Research

The work that has been conducted at the University of Texas has focused heavily on the fatigue

performance of various HMIP connection details. The work has included laboratory studies as

well computational investigations. These studies demonstrated the impact of the base plate

geometry and weld details on the fatigue performance. The research also showed that the

galvanizing process that is used to improve the corrosion performance leads to small cracks in the

welds between the shaft to base plate connection. These cracks have a significant impact on the

long term fatigue performance of the HMIP connections and are the primary focus of the current

stage of the research. The study on the cracking has focused on the behavior during galvanizing

that has been discussed by Kleineck (2011) and the work documented in this thesis that is

dedicated towards the behavior under wind loading in field conditions.

Page 23: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

8

2.1.1 Fatigue Performance - Phase I

The research in fatigue testing of High Mast Illumination Poles at the Univeristy of Texas at

Austin started in 2006. Specimens 14 ft long HMIP base sections designed with three different

baseplate to shaft connection details where investigated that are referred to as the socket detail,

the Texas detail, and the Wyoming detail. The connection details that were studied consist of the

following:

- The socket connection is realized inserting the shaft in the base plate hole, and then fillet

welded internally and externally as shown in Figure 2.1.

- The “Texas” detail is realized butting up the pole shaft to the baseplate, and welding

internally with a fillet weld. A weld access hole is therefore necessary in the base plate.

After the fillet weld is made, a full penetration weld is made between the shaft and the

base plate from the outside.

- The “Wyoming” connection detail still uses a full penetration weld, but a backing bar is

used on the inside of the shaft.

High Mast specimens with a 16 sided section were fatigue tested in the laboratory. The

specimens had a base shaft diameter of 24", and featured different connection details, variable

baseplate thickness, and different numbers of anchor bolts. The test setup consisted of two

specimens that were attached back to back to a stiff reaction box as shown in Figure 2.2. The

baseplates of the two specimens were connected at the middle of the setup to the reaction box,

that was connected to a hydraulic ram that reacted on a load frame anchored to a reaction floor.

The specimens were loaded at a frequency of 1.75 Hz in displacement control. The test was

interrupted when a reduction of 10% in the force to obtain the maximum displacement was

measured. The maximum displacement imposed was determined in such a way to cause a

nominal stress range at the base plate section of 12 ksi. Results from these tests showed that:

- The “Texas” and Wyoming details had good performance, satisfying AASHTO fatigue

category E’ requirements

- The socket connection detail had a poor fatigue performance, far below AASHTO

category E’

- Thicker base plate and higher number of bolts significantly increased the fatigue

performance of the specimens

Page 24: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

9

Fig. 2.1. Socket connection (top left), Texas connection (top right), Wyoming connection

(bottom) (Rios, 2007)

Page 25: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

10

Fig. 2.2. Load Testing Setup (Rios, 2007)

2.1.2 Fatigue Performance - Phase II

The second phase of the fatigue testing were documented by Stam, 2009. Specimens tested in

this phase of the study were fabricated and load tested with the same basic setup from the first

phase of testing. The specimens utilized slightly modified details from the previous

investigations. Some were fabricated with the use of an external collar that is referred to as a

ground sleeve. The ground sleeve, depicted in Figure 2.4, serves as a doubler plate around the

shaft.

Reaction

Box

1 DOF

ReactionSupport

2 DOF

ReactionSupport

55 kip MTS Ram

16 Sided,5/16" thick,24" diameterPole

Page 26: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

11

Fig. 2.3 Texas detail with External Collar (Stam, 2009)

Fig. 2.4 Wyoming detail with External Collar (Stam, 2009)

Page 27: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

12

Conclusions that were made from the second phase of testing consisted of the following:

- The Texas detail performance improved with External Collar and did not develop fatigue

cracking even after passing AASHTO category C

- The Socket connection with a 3" thick base plate satisfied the AASHTO category D

requirements

- Both the Texas and Wyoming details benefited from the reduction of baseplate inner

diameter

The second phase of the research also included a three-dimensional finite element model that was

developed in ABAQUS. The model was used to conduct a parametrical study to investigate the

variations in the Stress Concentration Factor (SCF). The SCF provides a measure of the

amplification of the nominal stresses at hotspots and provides an indication of the impact on the

fatigue performance. The SCF were obtained by normalizing hotspot stresses ( with the

nominal stresses ( :

(Eq. 2.1)

Fig. 2.5. Extrapolation of Hotspot Stress from Finite Element Model (Stam, 2009)

Page 28: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

13

From the FEM parametric study the following conclusions where made:

- Increasing the shaft diameter or the shaft thickness results in a lower hotspot stress

- Adding anchor rods results in a SCF reduction when the connection is flexible (no

external collar and thin base plate). The SCF does not change significantly with the

number of anchor bolts when the 3 in. thick base plate was used.

- A smaller diameter of the internal baseplate diameter results in a smaller SCF

- Adding an external collar to a socket detail can reduce the SCF by up to 40%

2.1.3 Fatigue Performance – Phase III

The specimens that were fabricated and tested in the third phase of the study featured a 12 sided

hollow section with a 33" base diameter. A 3" thick base plate with 12 anchor bolts was used and

the shaft wall thickness was 5/16". The connection details studied in this phase focused on the

Texas detail shown previously in Figure 2.1 and the Texas detail with external collar shown

previously in Figure 2.2a. These specimens match the connections of the HMIP’s that were

instrumented on the field on the present study and widely used throughout the state of Texas.

Therefore the results from these tests provide valuable information of the long term fatigue

performance of the Texas poles in service. These information is useful for the evaluation of the

fatigue performance that are presented in Chapter 5.

One of the specimens in the phase III testing was load tested prior to galvanization, in what is

called “black” state. This test was performed to investigate the effect of galvanization on the

fatigue performance of the HMIP. Concerns about the hot dip galvanizing process possibly

causing micro cracking above the weld toe had been raised by Koenigs (2003). Koenigs

investigated the fatigue performance of traffic signal mast arms, and noticed how galvanized

connections exhibited a reduced fatigue life as compared to the non-galvanized connections.

A “field” repair procedure and a “shop” repair procedure were also developed. These were

practical solutions for TxDOT to implement on in-service poles that were cracked. The shop

repair procedure requires an extended preparation of the welding with a grind depth of half of the

shaft thickness, and a length of two inches past the crack extent. The weld process is a flux core

arc weld (FCAW). The field repair procedure was developed for poles already in service that are

Page 29: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

14

found to be cracked. The surface preparation is similar to the one for shop repair, but the welding

is a shielded metal arc welding (SMAW). More information about the welding procedures is

provided by Pool (2010). Two HIMP specimens were repaired following each of these procedures

and load tested.

Table 2.1 shows the list of specimens from the phase III testing. The “black” specimen that was

load tested, but never failed, is not shown in the table.

Table 2.1 – Phase III HMIP fatigue testing specimens (Pool ,2010)

Figure 2.6 shows the results of the load testing performed in Phase III. Both the specimens with

the external collar connection did not fail to the test. The “shop” repaired specimen also never

failed.

Page 30: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

15

Fig. 2.6. Extrapolation of Hotspot Stress from Finite Element Model (Pool, 2010)

Conclusions that were reached from the Phase III study are as follows:

- The ultrasonic inspections performed on every pole before testing indicated that only

poles that had been galvanized had initial cracking.

- The likely source of the cracking was from the galvanizing process.

- Initial cracking significantly reduces the fatigue performance of the poles with the Texas

standard detail.

- The external collar provides improved fatigue performance of poles compared to poles

without the external collar.

The conclusions from this study reflect exactly what it has been seen in the field from the

ultrasonic inspections carried out by TxDOT on in-service HMIP around the state of Texas.

Page 31: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

16

2.2 Thermal investigation during Galvanization

Initial cracks can propagate in-service due to wind loads acting on the HMIP’s, significantly

reducing the fatigue life of the poles and increasing the potential for collapse. The concerns raised

by previous investigations about the initial cracking found after galvanizing, lead to a new phase

of the project. The thermal study during galvanization is aimed to determine the causes of the

initial cracking in HMIPs and is summarized by Kleineck (2011).

Fig. 2.7. HMIP specimen entering the zinc bath Fig. 2.8. FEM for Thermal Sudies

HMIPs are protected from corrosion by the hot galvanizing coating, which is formed by dipping

the pole into a molten zinc bath at a temperature of approximately 840 °F (450 °C). The thermal

stresses generated in the steel during this process were investigated with a three-dimensional

finite element model developed with the software ABAQUS. The model recreates the molten zinc

bath with a semi-space in which the steel element is subjected to heat transfer by conduction and

convection. Measurements at galvanizing plants during the process were taken to study the

behavior and also to provide validation data for the finite element model. The temperature of the

steel was monitored in multiple locations during the galvanization of HMIP specimens. The

measurements showed high temperature gradients between the baseplate and the shaft, that could

result in high stresses in the material. In addition to demonstrating the behavior of the HMIP’s

during galvanizing, the data was used to validate the finite element model of the galvanizing

process. Some of the modeling aspects that the data was used to help define were the coefficients

Page 32: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

17

for the heat transmission equations, the correct angle at which the HMIP is dipped in the zinc

bath, and the speed at which it is lowered in. Once the response from the model matched the

conditions measured on-site, a parametric study was conducted to determine what parameters

intervene in the likelihood of the HMIP to develop initial cracks.

Other important results were obtained by Ultrasonic Testing (UT) of the specimens. The

specimens were inspected using UT techniques before an after galvanizing. The specimens were

designed with similar geometries, but some had an external collar connection, while some did

not. In the current TxDOT design for poles that are 150 ft tall, with a design wind speed of 80

mph, the shaft wall is 5/16" thick. The results (Tab. 2.2.) show how the use of an external collar

reduces the amount of cracking caused by galvanizing.

Table 2.2. Ultrasonic Tests Results after Galvanization

Specimen Name

Shaft Wall

Thickness

(in)

Baseplate

Thickness

(in)

External

Collar

Bends

Cracked

(%)

(in)

33-3-12-TX-SG-C 5/16 3 100 27.6

33-3-12-TX-SG-SB 5/16 3 58 8.3

33-3-12-TXEC-SG-SA 5/16 3 √ 58 7.7

33-3-12-TXEC-SG-B 5/16 3 √ 25 8

33-3-12-TXEC-SG-SC 5/16 3 √ 17 5

These results will be discussed in chapter 5, in correlation with the data obtained from the field.

They will be used to define a fatigue performance of the in-sevice poles that are found to be

cracked.

2.3 Wind Effects on High Mast Illumination Poles

HMIP are slender tall structures, with multisided or circular hollow sections, which are mainly

subjected to wind loads. These loads are caused by buffeting from natural wind gusts and vortex

shedding. Vortex shedding involves the generation of pressure differences normal to the wind

direction, which cause the structure to oscillate transverse to the wind. The following sub-sections

provide a brief discussion about wind engineering concepts.

Page 33: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

18

2.3.1 Vortex Shedding

Vortex shedding occurs when alternating vortices are created on opposite sides of an object. The

creation of vortexes next to the object depends on the regime of flow, which is determined by the

Reynolds number :

(Eq. 2.2)

Where, is the wind velocity

is the diameter of the cross section

is the coefficient of kinematics fluid viscosity (1.56 x 10-4

ft2/sec for air)

Vortex shedding occurs with , and the regime is fully turbulent.

For vortex shedding to excite the structure, the vortex frequency has to be close to a natural

frequency of the structure. When this happens, the structure will be locked-in a resonant state

with the vortex shedding. The critical wind speed for vortex shedding is given by:

(Eq. 2.3)

Where, is the wind velocity for Lock-in

is the diameter of the cross section

is the Strouhal number, a constant that depends on the cross section shape.

To determine the dynamic load acting on the HMIP due to vortex shedding a model developed by

Scanlan can be used (Simiu and Scanlan, 1996). The model assumes that the vortex shedding

excitation is a sinusoidal process, and thus the forcing function is a harmonic.

The maximum displacement occurred during vortex shedding excitation depends on the mass and

damping properties of the structure. The Scruton number combines these two properties in a

unique quantity:

Page 34: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

19

(Eq. 2.4)

Where, is the mass per unit length

is the critical damping ratio

is the air density

is the diameter of the cross section

The following empirical relationship can be used to determine the maximum displacement under

vortex shedding excitation as a function of the Strouhal number and the Scruton number (Simiu

and Scanlan, 1996):

(Eq. 2.5)

Where, is the maximum displacement

is the Strouhal number

is the Scruton number

is the diameter of the cross section

2.3.2 Buffeting

The term buffeting indicates an unsteady loading caused by the wind on the structure, by flow

velocity fluctuations. The dynamic excitation of the structure due to buffeting increases with the

flow speed, and it is not limited to the natural frequencies of the structure as in vortex shedding.

2.3.3 Static Parameters

The static action of the wind under steady conditions can be simplified by a drag force acting

in the direction parallel to the wind, and a lift force acting transversally to the wind. The drag

coefficient and the lift coefficient need to be defined:

Page 35: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

20

(Eq. 2.6)

(Eq. 2.7)

Where, is the air density

is the wind velocity

is the diameter of the cross section

2.3.4 AASHTO Wind Design Specifications

The AASHTO Specifications (AASHTO, 2001) target a design life of 50 years for luminary

support structures.

The design wind pressure is computed as:

(Eq. 2.8)

Where, is the height exposure factor, which is equal to 1.37 for a height of 150 ft

is the gust effect factor, which shall be taken as a minimum of 1.14, or calculated

following more complex procedures presented in ASCE 7 (ASCE, 2005).

is the basic wind speed (80 mph, or 100 mph for the HMIPs considered in this study)

is the wind importance factor, which is taken as 1 for a design life of 50 years

is the drag coefficient, and for a dodecagonal ranges from 0.79 to 1.20 as

demonstrated in the graph shown in Fig.2.9.

Page 36: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

21

Fig. 2.9. Drag Coefficient for a Dodecagonal section as suggested by (AASHTO, 2006)

AASHTO only requires that the vortex shedding action is considered for non-tapered structures,

or structures having a taper (diameter diffenece over length) less than 0.14 in/ft. The taper of the

HMIP considered in this study is roughly 0.17 in/ft, thus vortex shedding design is not required.

The equivalent static wind pressure range for fatigue is then calculated as:

(Eq. 2.9)

Where, is the fatigue importance factor, which is taken as 1.0 for HMIPs installed on major

highways

is the drag coefficient, and for a dodecagonal ranges from 0.79 to 1.20 as shown in

Fig.2.9.

For non-tapered poles, the fatigue design is carried out considering the vortex shedding critical

wind speed for the first natural mode of the structure, which is from equation 2.3.

AASHTO suggests the Strouhal number to be taken as 0.18 for circular sections, and 0.15 for

multisided sections. The equivalent static pressure range to be used for the design of vortex

shedding induced loads is:

Page 37: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

22

(Eq. 2.10)

Where, is the vortex shedding critical wind speed for the first natural mode of the structure

is the critical damping ratio, which is conservatively taken as 0.005

and as defined in equation 2.9.

2.3.1 Iowa Study

An interesting study funded by the Iowa Department of Transportation (IDT) and Midwest

Transportation Consortium (MTC) was carried out at Iowa State University (Warpinski, 2006)

(MTC, 2007). Failures of HMIPs occurred in Iowa, as shown in Chapter 1. These failures

demonstrated a lack of knowledge in design specifications such as AASHTO. The study

developed with a field investigation on in-service poles, wind tunnel testing, and analytical

modeling of the induced wind loads, was aimed to reevaluate the current equations used for the

design of HMIPs.

Wind tunnel tests were performed on a 12-sided cylinder to calculate aerodynamic parameters.

The testing model was a 20" long wooden cylinder with a 12-sided cross section. The model was

secured transversally to the wind flow (Fig. 2.10) by mean four chains attached to coil springs.

Two leaf springs where also used to fix the model in the wind direction. To measure the drag

coefficient the coil springs where replaced by fixed supports, to restrain the model in the wind

direction. The force was measured by transducers at the extremities of the springs.

The Strouhal number was measured to be 0.2, which is different from the value specified in

AASHTO. The drag coefficient was measured at flow regimes with Re ranging from 2.5 x 104 to

2.25 x 105, and was found to vary from 1.4 to 1.6 (Fig.2.11.)

Page 38: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

23

Fig.2.10. Wind Tunnel Testing Setup from the Iowa Study (MTC, 2007)

Fig. 2.11. Measured Drag Coefficient for a 12-sided section (MTC, 2007)

Page 39: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

24

Field instrumentation and monitoring was performed on two HMIP in Northern Iowa, in high

wind locations. One of the HMIP’s was located in Sioux City and another in Mason City. The

HIMPs were 148 ft tall, with a 12-sided cross section, a 0.313 in thick shaft having base diameter

(measured flat side to flat side) of 28.5" and a top diameter of 8.77" (0.14in/ft taper). The base

plate was 1.25 in thick and was anchored to the foundation by six 2.25 in diameter bolts.

The instrumentation consisted of anemometers, accelerometers and strain gages. One of the poles

was instrumented with three anemometers distributed along the height of the structure and six

strain gages. For the second pole, one anemometer, four accelerometers and fourteen strain gages

were used.

The accelerometers allowed a measurement of the natural frequencies of the pole: 0.388Hz,

1.34Hz, 3.41Hz, 6.70 Hz, respectively for the first four modes of vibration (MTC, 2007).

The damping ratio was measured using a pluck test, which consists in exciting the structure with

an impulsive force and measuring the dynamic response. This was done pulling and releasing a

cable attached to the pole at a suitable height. Table 2.3 shows the measured modal frequencies of

the pole, obtained by a Fast Fourier Transform (FFT) of the recorded motion. Also shown are the

frequencies calculated analytically from a Finite Element Modeling (FEM) of the Structure

(Warpinski, 2007). The values of the damping ratio were found to be lower than the value

specified by AASHTO for design calculations.

Tab. 2.3. Modal Frequencies and Damping Ratios (MTC, 2007)

From the long term field monitoring it was observed that buffeting excitation was prevalently in

the first natural mode of the HMIP, while vortex shedding excitation was observed in the second

Page 40: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

25

mode of vibration. The critical wind velocity for vortex shedding in resonance with the second

mode was computed to be approximately 6.5 mph. Vortex shedding induced vibrations were

observed at a wind speeds ranging from 3 to 8mph. The strain gage measurements showed that

the maximum vortex shedding induced stress range exceeded the constant amplitude fatigue limit

of 2.6 ksi, as defined for AASHTO Category E' detail. The pole had a socket type connection that

is classified with the lowest fatigue performance, and has been observed to perform poorly in

fatigue (Stam, 2009). Figure 2.12 shows how the stress level in the wind transverse direction has

a peak at the vortex shedding critical velocity of approximately 6.5 mph.

Fig. 2.12. Measured Stress Range vs. Wind Speed from Iowa Study (MTC, 2007)

The highest stress range measured during the long term instrumentation was 12.6 ksi, at a height

of 5.75 ft from the base plate. High stresses were caused by buffeting in the first mode of

vibrations in presence of wind speeds higher than 20 mph. Since the cumulative frequency of

occurrence of wind stronger than 20 mph was lower than 5% on a period of 15 months, buffeting

was not believed to b contribute significantly to fatigue damage of the pole. Figure 2.12 shows

the average wind speed frequency of occurrence at the instrumented site.

Page 41: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

26

Fig. 2.13. Probability Distributions of Mean Wind Speed in Sioux City (MTC, 2007)

A mathematical dynamic model was then developed for the combined response to buffeting and

vortex shedding (Chang et al., 2009). Figure 2.13 shows the predicted stress range on the pole in

parallel and perpendicular to the wind direction.

Page 42: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

27

Fig. 2.13. HMIP analytical model response (Chang et al., 2009).

In Chapter 5, these results are discussed and compared with the results of the field measurements

obtained from studies of HMIP’s in Texas.

Page 43: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

28

Chapter 3 Filed Instrumentation

3.1 Objective of the Instrumentation

The objective of the field instrumentation was to collect long term stress and wind data from

multiple poles located around the state of Texas. Four Poles were instrumented and monitored

with a system that was accessible via a modem with a cellular link. Energy harvesting was

necessary to recharge the batteries that powered the system.

A total of five HMIPs were instrumented at four sites around the state of Texas as indicated in the

wind activity map shown in Figure 3.1. The activity map winds indicated on the map represent

the average wind speeds throughout the year. The instrumented locations include the following

sites:

1. Austin, Mo-Pac @ Parmer Ln.

2. El Paso, IH-10 @ Brown St.

3. El Paso, US-54 @ Hercules St.

4. Corpus Christi, IH-37 @ SH-286

5. Lubbock, US-62 @ SH-289

The site locations were selected in an attempt to capture the wide range of exposure conditions in

the state of Texas. The Austin location was primarily selected for convenience to evaluate the

system; however the location is also representative of the conditions found in several areas

around the state. The El Paso site included poles with significant cracking positioned in the

desert mountains of Texas. Lubbock is located in the Texas high plains. The relatively flat

region experiences significant wind throughout the year. The final location that was selected was

in Corpus Christi, which is an area representative of the Texas Coast. Figure 3.2 shows the base

section of the instrumented HMIP in Austin. Although the connection detail in the five poles

varied, the basic geometry of the poles were similar. The poles had twelve-sided shafts with a

base diameter of 33" and a height of 150 ft. The pole in Corpus Christi had the external collar

Page 44: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

29

connection between the shaft and base plate, while the other poles had the standard Texas

connection with the full penetration weld.

Fig. 3.1. Location of instrumented HMIPs on a wind activity map (www.whisperenergy.com)

Page 45: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

30

Fig. 3.2. Typical 12 Sided Pole instrumented (Austin, TX).

Page 46: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

31

3.2 Hardware

Fig. 3.3. Scheme of the Instrumentation: (1)High Mast Pole, (2)Anemometer, (3)Strain Gages,

(4)DAQ, (5)Wireless Modem, (6)Antenna, (7)Solar Panel, (8)Charge Controller, (9)Battery

The instrumentation that was used in the field included sensors, data acquisition, a power source,

and an energy harvesting system for recharging. Figure 3.3 shows a schematic representation of

the instrumentation setup. The sensors included strain gages and an anemometer that were

attached to the HMIP (1). The strain gages (3) provided a measure of the stresses induced from

wind acting on the pole, while the anemometer (2) provided the corresponding wind speed and

direction. The sensors were monitored with a data acquisition system (DAQ) (4) that was

programmed to record, process and store the data. A wireless modem (5) with an antenna (6) was

provided so that the system could be configured remotely and data could be downloaded. The

system was powered by a 12V battery (9), which was charged by a solar panel (7). A charge

controller (8) was provided to control the charge transfer between the solar panel, battery, and the

Page 47: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

32

data acquisition system. A plan view (Fig.3.4) of the setup in the El Paso Pole on IH-10 shows

the location of the anemometer brace and the four strain gauges on orthogonal faces named

N,S,E,W. A discussion of each of the components of the system is provided in the following

subsections.

Fig. 3.4. Plan view of the instrumentation on IH-10 (El Paso, TX)

Page 48: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

33

3.2.1 Stress Measurement

Foil strain gages with a 350 ohm resistance were installed on the HMIPs to measure wind-

induced stresses. Four gages were installed on orthogonal faces of the pole at a height of 76" (193

cm) from the base plate. The gages were installed above the access door to minimize the effect to

the stress distribution caused by the access panel opening on the shaft. The installation of the

gages required the removal of the galvanizing over a region that is approximately one square inch

in area at each gage location. A two part adhesive was used to install the gages. The adhesive

consists of a Cyanoacrylate component that is applied to the surface as well as a catalyst that is

applied to the back of the gage. After the gages were bonded to the pole shaft, microcrystalline

wax was installed along with a layer of silicon to protect the gages from weathering and moisture.

The galvanic protection to the shaft was restored using zinc spray paint following the removal of

the gages.

3.2.2 Wind Measurement

The speed and direction of the wind was monitored using the Wind Monitor Model 05103

anemometer manufactured by Young. The anemometer consists of a four-blade helicoid

propeller as shown in Figure 3.5. The anemometer provides DC voltage output ranging from 0 to

5 Volts for wind direction and AC voltage for the wind speed. An aluminum arm was fabricated

to support the anemometer as shown in Fig. 3.5 The aluminum support was attached to the pole at

a height of 35 ft (10.7 m), with stainless steel hose-clamp bands tightened around the pole. The

support inclination can be adjusted by means of four pointed screws that bear on the pole, so that

the anemometer can be aligned vertically. To minimize wind shielding of the anemometer by the

pole, the support was designed to be twice as long as the diameter of the pole. The anemometer

was oriented on the HMIP considering the prevalent wind activity in the area to further minimize

shielding in the strongest wind direction for each location.

Page 49: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

34

Fig. 3.5. Wind Monitor Model Young 05103, Helicoid Propeller Anemometer

Page 50: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

35

Fig. 3.6. Setting up the anemometer on

a pole (El Paso, TX)

Fig. 3.7. Solar panel and wireless antenna

mounted on a pole (El Paso, TX)

3.2.3 Data Acquisition and Communication

The CompactRIO 9024 from National Instruments (NI) was used for data acquisition in the field.

The system has an 800MHz processor, 512MB of RAM, and 4GB of Physical Memory. The

strain gages were connected to a NI 9237 module on the CompactRIO in a quarter-bridge

configuration whereas the voltage from the anemometer and battery were measured using a NI

9219 module. Measurements of the battery voltage provided valuable feedback on the power

conditions during the winter months when the recharging abilities of the solar panel were lowered

due to significant reductions in available daylight. An Airlink Raven X cellular modem from

Sierra Wireless was used so that the DAQ could be remotely accessed through the GPRS/EDGE

wireless network. The DAQ and the wireless modem were placed in a water resistant steel

enclosure inside the shaft, through the existing access port.

3.2.4 Powering the System

A battery recharged by a solar panel was used to power the data acquisition system since the high

voltage AC power supplied for the pole lights could not be used. To determine the nominal power

of the solar panel, a calculation of the power consumption of the system was performed. The total

power consumption, obtained by summing the maximum operating input power of the devices,

was estimated to be 24 Watts (W). Considering that the solar panel can only charge the battery

during the day, a factor of two was used. As such, the BP Solar BP350J, a 50 W nominal power

panel 33" (84 cm) long, 21" (53 cm) wide, and weighing 13.2 lb (6 kg), was chosen. A custom

Page 51: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

36

steel frame for the solar panel was fabricated to attach the panel to the HMIP as shown in Figure

3.7. The rack was clamped to the pole by means of two threaded bars. The panel was oriented

facing south at an inclination of 30°, which is the approximate latitude in Texas. The battery

chosen for the system is a flooded lead acid, deep cycle battery manufactured by Trojan. It has a

nominal capacity of 130Ah (amp-hours), a weight of 67 lb (30 kg) and dimensions of 13.25" x

6.75" x 9.75" (34 cm x 17 cm x 25 cm). The battery was placed in a plastic enclosure inside the

access door of the pole. The Morningstar Sunsaver 6 was the charge controller chosen, which has

a solar rating of 6 Amps and a max load of 10 Amps. The charge controller has a load disconnect

threshold of 11.5 Volts and a load reconnect of 12.6 Volts, and ensures that the battery is fully

charged before it starts powering the system again.

3.3 Software

National Instruments LabVIEW was used to create a program for capturing, processing, and

recording the data. The program captures data at 50 Hz (samples per second). The software

incorporated several features including recording raw data as well as conducting a rainflow

analysis for fatigue evaluation. The following subsections provide an overview of the features of

the data acquisition program.

3.3.1 Rainflow Counting of Stress Cycles

A fatigue life assessment is typically based on the cumulative damage theory as presented in

Miner (1945). The theory allows a complex stress history to be represented as a series of simple

constant amplitude cycles. The number of stress cycles are usually estimated using the Rainflow

Counting method which keeps track of the number of cycles within a preset series of stress

ranges. In the rainflow method, the wide range of expected variations in the applied stress cycles

is divided into a series of smaller increments. For example, if the cyclic stresses are expected to

vary from 0 ksi to 10 ksi, a uniform increment in stress range can be selected such as 0.25 ksi.

The stress range “bin sizes” would then be 0.25 ksi, 0.5 ksi, 0.75 ksi, etc. The Rainflow counting

algorithm then keeps track of the number stress cycles within a given stress range over time.

Figure 3.8 shows a simplified representation of the a variable stress history and the its Rainflow

Counting.

Page 52: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

37

In the literature many rainflow algorithms have been presented. The rainflow counting used in the

program was presented in Downing and Socie (1982). By default, the program performs the

rainflow analysis for each of the four strain gages every 30 minutes, using bin sizes of 5

microstrain and neglecting cycles smaller than two microstrain (equivalent to 0.06ksi)

Fig. 3.8. Stress History and Rainflow Counting example

3.3.2 Event Capture

In addition to gathering rainflow stress range data, the program was also written to be triggered to

gather raw data for a specified amount of time (usually 5 minutes) in the event that the wind

exceeded specified threshold values. The raw data included strain values along with wind speed

and direction. This data is useful to study the dynamic behavior of the pole as related to potential

vortex shedding effects. After collecting a library of recordings, approximate relationships

between wind speed and maximum stress can also be defined.

3.3.3 Voltage Measurement

Additional data that was monitored on the data acquisition system was the battery voltage, which

provided valuable information on the performance of the system in terms of power. The

instrumentation was installed during the summer months when there was significant sun exposure

Page 53: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

38

to provide energy to the solar panels that recharged the battery. During the summer and early fall

periods, solar exposure was sufficient to maintain the battery charge and keep the system running

continuously. However, as the sunlight intensity diminished during the winter months, the solar

panel could not recharge the battery fully. In addition, the colder temperatures in the winter

reduced the overall capacity of the battery. These factors caused the system to turn off more

frequently. If weather forecasts predicted storms (with higher winds), the readings of the battery

voltage allowed the system to be turned off so that the battery could charge properly so that storm

data was not missed. As expected, the performance of the system improved during the spring as

the days became longer and the solar exposure increased.

3.3.4 Remote Access and Control

Fig.3.9 is a screenshot of the program interface for the remote control of the DAQ. The plot on

the top right shows the real time values of the strain (in millistrain) measured by the four strain

gages, named N-S-W-E. The Rainflow Information is provided in the lower right hand quadrant

below that graph. The input parameters for the rainflow sampling consists of the length of over

which the rainflow analyses is performed, the size of the bins in microstrain, the number of bins,

and the minimum amplitude of strain cycles that should be considered in the rainflow analysis.

The live readings from the anemometer are displayed in the bottom left corner. On the bottom

right corner the control parameters for the triggering program used for the event capture. The

parameters that can be specified include the wind speed that defines the triggering point for the

event capture and also the duration of the event to be recorded.

Page 54: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

39

Fig. 3.9. Screenshot of the Instrumentation Program

Page 55: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

40

Chapter 4 Background Information for Field Studies

4.1 Introduction

The data gathered from in-situ HMIP’s provides valuable feedback on the actual performance of

the poles. However, because of the wide variability of both pole geometry and site

characteristics, the researchers considered information gathered from inspections and a database

of weather history to select poles for instrumentation. This chapter provides an overview of the

data that was considered in selecting candidate poles for instrumentation. The chapter has been

divided into three sections. Following this introduction, an overview of data gathered from

TxDOT inspections is provided. The data focuses on ultrasonic inspections that were conducted

for crack identification and crack growth monitoring. The final section of the chapter provides an

overview of the historical weather data for the candidate sites that were selected for field

monitoring.

4.2 TxDOT Ultrasonic Inspections

In an effort to determine the severity and range of damage on existing HMIP’s around the state,

TxDOT initiated an effort to inspect the state’s HMIP inventory. The main instrument used for

the inspection was Ultrasonic Testing (UT) in which a pulse wave is introduced into the HMIP

material through a transducer that also serves as the receiver as the signal is reflected back to the

device. Cracks or defects in the material affect the amount of sound that is reflected back to the

transducer. An experienced technician can use the measurements to identify and measure cracks

in the material or welds. In common practice, transducers with a 45° angle working in a 2.0 to

2.5 MHz range are typically used. However because the initial galvanizing cracks were relatively

small and difficult to detect using standard procedures, the UT inspections on the HMIPs were

performed using a 3.5MHz, 70° transducer with a ½" diameter. The higher ultrasonic frequency

increases the sensitivity of the measurements, and the interpretation of the results can be

dependent on the experience of the technician.

Another non-destructive inspection method that was tested is magnetic particle testing in which a

magnetic field is introduced into the region and iron filings are used to identify discontinuities

(cracks) (Pool, 2010).

Page 56: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

41

4.2.1 UT Results on Texas In-Service HMIPs

The data that was obtained in the UT tests on in-service poles provided insight into the condition

of the states HMIP inventory. The results from these tests along with variations in the sight

conditions around the state provided valuable data into the selection of poles that should be

instrumented for long term monitoring. The UT results for 12-sided poles from around the state

are summarized in Table 4.1. The inventory of poles were divided into two main groups

consisting of 80 mph designs and 100 mph designs. Four different pole heights existed within

each design wind speed: 100, 125, 150, and 175. Two different details were used between the

shaft and the baseplate. Details designated as “TX” represent a connection with a full

penetration weld with no external collar as outlined in Chapter 2, while those designated as “EC”

had an external collar.

Table 4.1. TxDOT Ultrasonic Testing Results

12 (sides)

80 (mph) 100 (mph)

100' 125' 150' 175' 100' 125' 150' 175'

TX EC TX EC TX EC TX EC TX EC TX EC TX EC TX EC

#UT 7 3 1 3 3 36 8 17 9 0 3 26 4 17 8 23

#Cr 4 2 1 1 3 36 3 9 2 0 0 0 1 5 6 2

%Cr 57 67 100 33 100 100 38 53 22 0 0 0 25 29 75 9

D/T 79 79 90 90 104 104 97 97 68 68 66 66 77 77 75 75

In the table the row labeled #UT is the number of HMIPs that were inspected while the row with

#Cr represents the number of poles in which cracks were found. The %Cr is the simply the

percentage of the poles inspected in which cracks were found, while the row labeled D/T provides

the ratio of the diameter at the base of the shaft to the shaft thickness. Previous observations have

been made in the correlation between the D/T ratio and the likelihood for the HMIP to have

cracks. This correlation can be demonstrated in the graph of the percentage of cracked poles

versus the D/T ratio as shown in Figure 4.1 (Pool, 2010). The detail with the highest rate of

cracking is the 80mph 150ft tall HMIP (100% cracking rate), which has also proved to have a

poor performance in terms of fatigue. A high ratio between the diameter and the thickness could

Page 57: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

42

indicate a potentially worse performance of the pole during the galvanization process. For a large

diameter with a thin shaft it would appear that the temperature difference during the dipping

phase of the pole in the molten zinc might create larger gradients between two sides of the pole,

thus leading to higher thermal stress generation.

Fig. 4.1. TxDOT Ultrasonic Testing Results

4.2.2 Cracking on Instrumented HMIPs

Based upon the geographic location and environmental exposure, poles were selected in Austin,

Corpus Christi, El Paso, and Lubbock for field instrumentation. The selected locations were well

representative of the range of environmental exposure that HMIPs might be subjected to around

the state. Ultrasonic testing inspections were performed on the specific poles so that the cracking

condition of the poles was also known prior to instrumentation. The UT results are summarized

in Table 4.2. The measured crack length in inches for each of the 12 bends is indicated in the

table. A check mark for a given bend indicates that no crack was found. Although the poles in

Lubbock and Corpus Christi had very little cracking; two poles in El Paso were selected since

they had significant cracking. The El Paso poles were actually inspected on multiple occasions

due to concerns about the cracks growing over time. One of the poles in El Paso was located on

Page 58: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

43

IH 10 while the other pole was on US 54. The latest inspection the HMIP on IH10 in El Paso was

found to have 11 out of 12 bends cracked, with a total crack length (summation of all the cracks

lengths) equal to 32" which is almost 1/3 of the overall circumference of the base shaft.

Considering three consecutive inspections of the IH10 pole in different times of the year, the

amount of cracking observed was found to be inconsistent from the first to the second inspection.

Since it is not physically possible that the cracking decreases with time, it is believed that this

inconsistency was likely due to the following reasons:

- Ultrasonic Testing is a sensitive and operator-dependable measurement, especially when

detecting microcracks such as those present in the welds. It is therefore possible that the

accuracy of the readings from the technician could vary from one day to the other.

- It has been observed on small sections of the HMIP baseplate to shaft connection cut

after fatigue testing (Figure 1.3), that cracks open and become more visible when the

specimens are heated to a higher temperature. It is therefore possible that the ambient

temperature had an effect, making the cracks more evident on the first inspection

(conducted in August 2008), and making the cracks close and become less detectable in

the second inspection (conducted in April 2009).

- The heat transmitted to the pole by sun radiation also causes a thermal gradient on the

HMIP. The sunny side can be several degrees warmer than the shaded side. This gradient

leads to a differential thermal expansion of the two sides of the pole, which causes the

HMIP to bend as depicted in Figure 4.2. This effect is visible by looking up at the poles

on sunny days. The second order moment generated by the bending may open the cracks

on the sunny side, and close the cracks on the shaded side. Further investigation was

carried out with a finite element model that is discussed in the next chapter. The model

was used to quantify the bending moment at the base of the pole generated by this effect.

If the reason for the inconsistency of the UT measurements is a thermal mechanism, then the

higher estimate of the crack length would be the most correct. If the difference is due to

variability in the reading by the operator, an average of the multiple readings would be the most

reasonable estimate. It would be interesting for TxDOT to report the ambient temperature and the

shaft temperature on the sunny side for every UT inspection performed. This would be helpful in

determining the nature of this change in measurements.

Page 59: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

44

Table 4.2. UT results for the field-instrumented HMIPs

HMIP 1 2 3 4 5 6 7 8 9 10 11 12 Date

% o

f B

en

ds

cra

ck

ed

TO

T C

rac

k

len

gth

[in

]

El Paso US-54 2.0 1.2 1.0 0.5 √ 2.0 2.1 2.0 1.1 1.0 √ 2.7 4/8/2009 83 15.6

" 2.0 1.2 1.0 0.5 √ 2.0 2.1 2.0 1.2 1.0 √ 2.7 6/15/2010 83 15.7

El Paso IH-10 √ √ 2.6 * 2.3 2.2 2.7 3.2 2.7 2.7 2.7 2.7 8/18/2008 75 21.1

" √ √ 2.9 √ 3.2 3.2 3.0 √ 3.0 √ √ 3.8 4/7/2009 67 15.3

" 0.8 √ 3.7 1.0 3.3 3.2 2.7 4.0 3.3 3.0 2.9 4.0 6/14/2010 92 31.9

Lubbock1 √ √ √ √ √ √ √ √ √ 0.9 √ √ - 8 0.9

Corpus Christi2 √ √ √ √ √ √ √ √ √ √ √ √ - 0 0

Fig. 4.2. Schematic representation of differential heating due to sun radiation heating of the

HMIP

1 100 mph wind speed design with external collar 2 100 mph wind speed design with full penetration weld (Texas detail) * small spot crack

Page 60: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

45

4.2.3 UT results of the HMIP specimens load tested in previous research phases

Ultrasonic tests were also performed on HMIPs specimens before being load tested at Ferguson

Structural Engineering Laboratory in the previous phases of the project. Table 4.3 shows the

results from the UT inspections on the specimens. The ultrasonic inspections were carried out by

a TxDOT technician, with the same procedure used for the in-service HMIPs. All the poles

reported in the table are 12 sided, 150ft tall designed for 80mph wind speed. The difference is

that some of the poles had the external collar connections, while others did not. On the right end

column the number of cycles to failure from the load test performed at a stress range of 12ksi is

shown. The specimen that showed the worst performance had a Texas standard shaft-to-baseplate

connection, with the full penetration weld and no external collar. It seems that within HMIPs

featuring the same connection detail the amount of initial cracking has an impact on the fatigue

performance. As expected, the specimen that had larger initial cracks showed the poorest

performance.

Table 4.3. UT results on HMIP specimens before load testing

HMIP #, Station External Collar

Galvanized % of Bends

cracked TOT Crack length [in]

Cycles to Failure @ 12ksi

33-3-12-TXEC-SG-A Yes Yes 17 5.00 No failure

33-3-12-TXEC-SG-B Yes Yes 25 8.00 No failure

33-3-12-TX-SG-A No Yes 83 11.00 81326

33-3-12-TX-VG-A No Yes 25 5.00 358228

33-3-12-TX-VG-B No Yes 33 4.63 358228

33-3-12-TX-SG-B No No 0 0 No failure

The crack data from these previous experiments combined with the number of fatigue cycles to

failure are useful for predicting the fatigue performance for the instrumented HMIPs. Indeed, if

there is a correlation between the initial cracking conditions and the fatigue performance,

comparing the UT results for the HMIP instrumented to the load tested specimens it will give an

idea of expected fatigue performance. The HMIP on US54 in El Paso, for instance, had initial

cracks prior to the field tests that were similar to specimen 33-3-12-TX-SG-A, highlighted in bold

in the table. The laboratory specimen had 83% of the bends cracked and a total crack length of

11", which is a little less than the length of approximately 15" measured on the El Paso US 54

Page 61: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

46

HMIP. It could be then assumed that the pole would have shown a similar performance under the

same loading conditions. The pole on IH 10 had an even larger amount of cracking at the shaft to

baseplate connection, with cracks on 92% of the shaft bends and a total crack length of about 30".

This HMIP was removed from service in July of 2011, and will be fatigue tested at Ferguson

Structural Engineering Lab using the same setup used for the previous investigations. This test

will demonstrate the impact of severe cracking on the fatigue performance of the pole.

4.3 Wind Historical Records

There are a number of organizations that archive meteorological data including statistics on the

wind historical records. These organizations make the data available to public. The following

sources provide wind historical data that can be used for various sites:

- National Climatic Data Center (NCDC), a US governative agency that offers the world’s

widest archive of meteorological data.

- Weatherunderground, a private company that owns 19000 weather stations across the US

and makes its information accessible through its website www.weatherunderground.com.

- Texas Commission on Environmental Quality (TCEQ), the environmental agency for the

state of Texas

The wind historical records were helpful in selecting pole locations for instrumentation. One of

the major goals of the field instrumentation was to gain a measure of wind induced stresses in the

pole. Higher wind speeds will create larger stresses in the poles. In addition, the propensity for

vortex shedding and the various dynamic modes are sensitive wind speed. Therefore, reviewing

the historical wind records were important in selecting sites that might result in significant wind

exposure for the instrumented HMIPs.

Tables 4.4 and 4.5 show the average and the maximum wind speed measured in the past in the

three cities that were selected for HMIP instrumentation. The right end column shows the time

length of the reference data. The average yearly wind speed in Lubbock and Corpus Christi is

around 12 mph, while in El Paso the average wind speed is almost 9 mph. On the other hand El

Paso had the highest measured gust of 86mph compared to 85 mph for Lubbock and 67 mph for

Corpus Christi.

Page 62: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

47

Table 4.4. Average Wind speeds in the Instrumented HMIP locations (NCDC)

JAN

FEB

MA

R

AP

R

MA

Y

JUN

JUL

AU

G

SEP

OC

T

NO

V

DEC

V a

vera

ge

Re

cord

ed

Pe

rio

d [

yr]

CORPUS CHRISTI, TX 12 12.9 14 14.3 12.8 11.7 11.5 11 10.4 10.4 11.7 11.6 12 60

EL PASO, TX 8.3 9.1 10.9 11 10.3 9.3 8.3 7.7 7.6 7.5 8 7.9 8.8 60

LUBBOCK, TX 12 13.2 14.6 14.7 14.2 13.6 11.4 10.1 10.5 11.2 11.7 11.8 12.4 53

Table 4.5. Maximum gusts speeds in the HMIP locations (NCDC)

JAN

FEB

MA

R

AP

R

MA

Y

JUN

JUL

AU

G

SEP

OC

T

NO

V

DEC

V M

AX

Re

cord

ed

Pe

rio

d [

yr]

CORPUS CHRISTI, TX 52 60 54 67 60 61 49 48 61 53 60 54 67 60

EL PASO, TX 86 60 69 66 55 69 63 48 62 53 54 61 86 60

LUBBOCK, TX 59 64 77 71 74 85 72 59 58 52 63 64 85 53

Page 63: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

48

Fig.4.3. Rate of occurrence of the daily average wind speed at the HMIP Locations

Figure 4.3 is a plot of the rate of occurrence of the daily average wind speeds in the last 3 years.

For El Paso the wind speed with the highest rate of occurrence is from 6 to 8 mph, while for

Lubbock and Corpus Christi the wind speed with the highest rate is from 8 to 10 mph. Lubbock

and Corpus Christi are more likely to have sustained wind (10 to 20 mph) on an average basis.

Another useful source of information for the instrumentation layout are “wind roses” that consist

of radial plots that show the probability distribution of the wind in terms of speed and direction.

This information can be used to position the anemometer on the pole to minimize wind shielding

in the direction that is most likely to have the prevailing wind. Figure 4.4 shows that Corpus

Christi has a prevalent wind action happening in the South-East direction, while Lubbock has a

prevalent wind happening from the South direction. The El Paso plot shows wind activity in

almost all the directions with peaks at North and South-East. It can also be seen that in El Paso

strong wind events (higher than 20mph) from the West are prevalent.

Page 64: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

49

Fig. 4.4.Yearly wind rose for Corpus Christi (Texas Commission on Environmental Quality)

Page 65: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

50

Fig. 4.5. Yearly wind rose for El Paso (Texas Commission on Environmental Quality)

Page 66: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

51

Fig. 4.6. Yearly wind rose for Lubbock (Texas Commission on Environmental Quality)

Page 67: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

52

4.4 Summary

Significant background information was used for identifying the poles that should be

instrumented for field monitoring. The data included ultrasonic testing that identified cracks in

the welds between the shaft and the base plate of the HMIP as well as wind historical records.

The wind historical records provided valuable data on the average wind speeds in each site as

well as the maximum gusts that have occurred at each site. The data also provided an indication

of the prevailing wind direction for use in positioning the anemometer to avoid wind shielding

from the HMIP. The UT results from the poles from around the state provided insight into the

pole geometries that had the largest probability of cracking from the galvanizing. For each pole

that was instrumented, the UT results showed the degree of cracking that was present. While this

data was not necessarily used in the selection of all of the instrumented poles, in the case of the El

Paso pole on I10, the data was used since it was desired to see if the damage was escalating based

upon severe wind exposure.

Page 68: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

53

Chapter 5 Data Analysis

5.1 Supporting Finite Element Modeling (FEM)

In an effort to complement the field measurements, two linear elastic models were created to

support the data analysis. The models include a simple cantilever beam model and a shell element

model. The beam model was developed to simulate the dynamic behavior of the high mast pole,

while the shell element model of the pole base section was put together to investigate the

influence of the anchor bolts on the stress distribution. Both the models were developed using the

structural analysis commercial software SAP2000 Nonlinear. The following two subsections

provide an overview of the computer models as well as the results that were used to assist in the

data analysis from the field work.

5.1.1 Beam FEM

To match the taper of the HMIPs, the beam model was composed of non-prismatic beam

elements, with a point mass on the top to model the lighting apparatus. The total weight of the

HMIP is approximately 14300 lbs., including 12500 lbs. for the pole and 1800 lbs. for the

lighting apparatus at the top. The model is restrained from translation and rotation at the ground

to simulate the fixed end condition of the cantilever poles. The model was used to conduct a

modal analysis to determine the natural period for the first three modes of the structure as shown

in Fig.5.1.

The bending moment diagram along the length of the pole was determined for each natural mode

as shown in Fig.5.2. This information was used to determine the ratio between the stress

measured in the field using the strain gages located 76" (193 cm) from the base and the nominal

stress at the base. A node in the finite element model was positioned at the strain gage location to

provide the value of the moment at the strain gage location. The FEM moment shape

corresponding to the respective first, second, and third modes shows that the stress at the base of

the HMIP are 5%, 11%, and 23% greater than the stress at the strain gage location. These values

were used to adjust the strain gage measurements to determine the nominal stress at the base.

Page 69: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

54

Fig.5.1. Natural Periods of the HMIP from the Finite Element Model

Fig.5.2. Bending moment shapes for the natural modes of the HMIP

Page 70: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

55

The natural frequencies calculated by the modal analysis were used to determine the wind speeds

that could potentially cause vortex shedding. Recalling the equation presented in section 2.3:

(Eq.5.1)

Where, is the wind velocity for vortex shedding lock-in

is the diameter of the cross section

is the Strouhal number,

The above equation is intended for prismatic sections. To use the expression on the tapered

HMIP, an the diameter of the HMIP at 2/3 of the total height was used to approximate an

equivalent prismatic section, which is also the location that an equivalent wind push was located.

For the poles being studied, the diameter at 2/3 of the total height was 16". The Strouhal number

was taken as 0.15, as suggested by AASHTO for a multisided section. Table 5.1 shows the values

of the vortex shedding critical wind speed calculated for the first three natural modes.

Table 5.1. HMIP calculated natural frequencies and periods

I MODE II MODE III MODE

[Hz] 0.28 1.20 3.57

[mph] 1.7 7.4 21.6

As noted in the discussion of the data acquisition system, event capturing was used to gather data

for very high wind speeds or behavior for other conditions. Event capturing was used to

investigate the behavior of the pole in presence of the wind speeds shown in Table 5.1, to

understand if there is any vortex shedding effect on the HMIPs. The frequency analysis of the

recorded strain was used to evaluate the accuracy of the model in predicting the natural period of

the structure. This is discussed in subsection 5.2.2.

The model was also used to evaluate the deformation of the HMIP under a temperature gradient

as explained in section 4.2.2. To determine a suitable value for the temperature gradient between

the sunny and the shaded side of the pole, measurements were taken on a pole shaft throughout

Page 71: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

56

the day using an infrared thermometer. The largest temperature difference recorded was 30F

which occurred on on a spring day at approximately 10AM. The ambient temperature at this time

was 60 F, and the sun radiation heated the steel surface on the sunny side to above 90F. Since the

temperature difference was measured close to the bottom section of the pole, the gradient was

calculated by dividing the gradient by the bottom shaft diameter of 33". The resulting temperature

gradient of 0.9 F/in was then applied to the frame elements of the model, and a second- (P-Delta)

order large displacement analysis was conducted. The coefficient of linear thermal expansion

used was set to .

The analysis results show a displacement of 10" at the top of the pole. The value of the second-

order moment generated by the structural weight at the base of the pole was computed to be 43 k-

in. Whether such a moment could affect the results of the UT inspections will be confirmed in

the next phase of the project. The HMIP in El Paso on IH10 has been removed from service and

will be fatigue tested at Ferguson Structural Engineering Lab. It will be easy to apply such a

bending moment on the specimen once set up in the loading frame. Multiple UT inspections can

then be performed with the moment applied in the opposite direction to determine the impact on

the readings.

Page 72: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

57

Fig.5.3. Displacement shape (left) and Moment diagram (right) due to sun temperature gradient

and P-delta effects

Page 73: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

58

5.1.2 Shell FEM

In some in HMIP pole installations, the anchor bolts are not properly tightened or the nuts are

even missing. The impact of the improper anchor installations on the stress conditions at the base

is not well understood. A three-dimensional finite element analytical model was therefore created

to evaluate the impact on the stress concentration factor (SCF) with differing support conditions.

The model represents a 10 ft (3 m) long section with a 3" (76 mm) thick base-plate that was

restrained at the base with 10 anchor bolts. The access door opening was also modeled to obtain

an indication of influence on the stress distribution. The section was subjected to a concentrated

bending moment applied at the top. The moment is applied through 12 rigid beam elements

connected to each shaft bend as depicted in Fig.5.4.

Multiple simulations were run on the model taking out some supports in an effort to model the

lack of anchor bolts on the pole behavior. The value of the stress at the most critical bend of the

shaft was then compared to the value from the analysis with the full anchor bolt connection. The

value of the stress to be compared was the average stress in the shell element in the vertical

direction (S22) at the most critical bend. Table 5.1 shows the results of this parametric study.

Page 74: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

59

Fig.5.4. Shell Finite Element Model

Page 75: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

60

Fig.5.5. Shell stresses S22 [ksi]

Table 5.2. Influence of anchor bolts on fatigue life

MISSING

SUPPORTS

QT, (#ID)

STRESS

(ksi)

STRESS

RATIO

FATIGUE

LIFE

REDUCTION

- , - 4.55 1.0 1.0

1, (#5) 4.76 1.05 1.1

2, (#5 #7) 5.24 1.15 1.5

3. (#4 #5 #7) 5.54 1.22 1.8

Page 76: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

61

The values on the right end column of the table represent a measure of how the fatigue life of the

HMIP was affected with a reduction in the number of anchor bolts. This factor was calculated

assuming that the fatigue life is proportional to the stress range to the power of three, as stated by

Miner (1945). Miner’s theory is discussed in Section 5.2.

The positive effect of a higher number of anchor rods on the fatigue life of the HMIP had been

pointed out in previous studies (Warpinski, 2006, Rios, 2007, and Stam, 2009). Stam (Stam,

2009) also investigated this problem in a three-dimensional FEM study. He investigated

anchorages with 8, 10 and 12 rods distributed symmetrically, concluding the higher number of

bolts not to be significantly improving the fatigue performance when the base plate was thick,

such as the 3" plate present in the HMIPs considered in this work. However, in the application

discussed in this section, the anchor bolts were not symmetrically placed around the HMIP,

thereby simulating a condition of improper tightening or missing nuts on the anchor bolts. Such a

situation, creates a different stress distribution, with higher concentrations at the hotspots.

The problem has also been investigated in a field study part of the research reported by

Warpinski, 2006. During the pluck test (described in Section 2.3.1), some anchor bolts were

loosened to simulate improper installation. With just one nut loosened, an increase of stress was

noticed in the vicinity of the loosened anchor point. It needs to be pointed out that the base plate

thickness of the HMIP under consideration was only 1.25" thick, which makes the stress

distribution more susceptible to change with the number of anchor points.

5.2 Event Capture Analysis

This section focuses on the data gathered from the event capture feature of the data acquisition

system that was discussed in Section 3.3.2. A triggering program was used to capture real time

records of strain and the corresponding wind speed and direction at a frequency of 50Hz. The

recording started when the wind speed exceeded a specified threshold. A library of 5 minute long

events was captured at wind speeds ranging from 0 to 50 mph during the period of

instrumentation. Many records were taken at wind speeds that could potentially cause vortex

shedding, in the 7mph range and 20mph range as discussed in the previous section.

Page 77: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

62

5.2.1 Data Observation

Figure 5.6 shows a graph of a 10 second record for an event with a wind speed ranging from 30 to

50mph. The wind speed shown is averaged over a three second period, while the stress is

averaged over a one second period. The average wind direction during the recording was 237

degrees (W-S-W). The plot shows the stress calculated on the east side of the pole, which is a

measure of the pole bending in the East-West direction. From the recording it can be seen that the

HMIP is shaking with a period that is approximately 3.5 seconds, with a 7 ksi stress range. This

indicates that, at high wind speeds, the first mode is the primary mode that is excited, and that the

wind effect is prevalently buffeting, exciting the pole in the direction of the wind rather than

transversally.

Fig. 5.6. Event recorded on the El Paso US-54 pole at a 50mph wind speed

Page 78: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

63

Fig. 5.7. Event recorded on the El Paso US-54 pole at a 7 mph wind speed direction S-S-W

Figure 5.7 shows a graph of a 10 second event record for a 7 mph wind speed. The average wind

direction during the recording was 200 degrees (S-S-W). The wind speed is averaged over a three

second period, while the stress is averaged above 0.2 seconds. From the recording it can be seen

that the HMIP is shaking with a period that is approximately 0.87 seconds, with a 0.2 ksi stress

range. This is an indication of vortex shedding exciting the second mode of the structure. This is

consistent with the observations from Warpinski (2006) that was discussed in Section 2.4. The

difference is in the amount of stress generated in the pole, which in this case remains below 1.5

ksi compared to those reported by Warpinski of values as large as 4 ksi for the vortex-shedding

second mode induced vibrations (MTC, 2007).

Figure 5.8 shows an event recorded on the Lubbock pole with a 25 mph wind and a direction 300

degrees (W-N-W). The plot shows the stress measured by the North strain gage, which is

measuring mostly cross wind swinging. The strain contains first mode vibrations mixed with a

small number of third mode cycles. The period was approximately 0.28 seconds, and had

Page 79: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

64

amplitudes smaller than 0.2 ksi. The third mode vortex shedding induced vibration in the

measurements in this study cause small stress variations and are judged not to contribute

significantly in the fatigue damage of the pole. This is consistent with the findings from previous

studies (MTC, 2007).

Fig. 5.8. Event recorded on the Lubbock pole at a 25 mph wind speed, direction W-N-W

From the observation of the data it can be concluded that the first and second modes are mostly

involved in the excitation of the structure. The first mode excitation occurs at a high wind speed

that can cause considerable stress in the structure, but are less likely to occur on an average basis.

The second mode induced vibrations have a lower stress range and happen at a lower wind speed

and are therefore more common.

Page 80: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

65

Considering the wind activity at the locations where the HMIPs were instrumented, it has been

shown that the wind speed with the highest probability of occurrence is approximately 8 mph for

El Paso and 11 mph for Corpus Christi and Lubbock. Observing an event with an average wind

speed close to 10 mph provides an indication of the average dynamic response of the HMIP.

From Figure 5.9 it can be seen that the response is a combination of first and second modes, with

stress ranges of roughly 0.2 ksi for the first mode and less than 0.1 ksi for the second mode.

Fig. 5.9. Event recorded on the Lubbock pole at a 10 mph wind speed, direction S-W

Further evaluation will be done to determine how much each of the first and second modes

contribute to the fatigue damage of the pole.

Page 81: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

66

5.2.2 Frequency Analysis

This section presents an analysis of the event capture data based on frequency content. The

analytical tools used for this analysis are mainly the Fast Fourier Transform (FFT), and the Root

Mean Square (RMS).

The FFT indicates an algorithm that produces a Discrete Fourier Transform (DFT) of a generic

signal that is the representation of the signal itself in a frequency domain rather than in the time

domain. The FFT performs the calculation with a smaller number of operations than the classic

DFT, making it suitable for a computational use. More information about the FFT procedure can

be found in the literature [e.g. Clough & Penzien, 1994; Paz, 1997; Lynn & Fuerst, 1998].

The root mean square (RMS) of the FFT allows the representation of the acquired time history

signal in the frequency domain but keeps the same physical quantity of the signal on the Y-axis.

For example, if the strain-time history is investigated, strain is graphed on the Y-axis and

frequency is graphed on the X-axis while if stress-time history is investigated, stress is graphed

on the Y-axis and frequency is graphed on the X-axis. The general expression of the cumulative

RMS is:

(Eq.5.2)

The Fourier spectrum for a typical strain record is shown in Figure 5.10. The frequency is

reported on the horizontal axis, and the Fourier amplitude is graphed on the vertical axis. The

peaks correspond to the first three modes of vibration of the HMIP. Comparisons between the

frequencies obtained from the FFT of the strain record and the frequencies calculated with the

finite element model are shown in Table 5.3. The comparison shows that the FEM was very

accurate for the first two modes with less than 2% difference, while the third mode had a 12%

difference.

Page 82: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

67

Fig. 5.10. Fourier Spectrum of a strain gage record

Table 5.3. Comparison between the FEM frequencies and FFT frequencies

FEM

[Hz]

FFT

[Hz]

Difference

[%]

I MODE 0.28 0.28 -

II MODE 1.20 1.22 2

III MODE 3.57 3.20 12

It is also interesting to note that the natural frequencies are similar to the ones calculated in the

Iowa Study for the poles in Sioux City (see section 2.5). Despite the different design, the poles

have similar dynamic properties, with frequencies differing only to 10% from one to the other.

The Iowa poles had indeed the same height (150 ft), the same cross sectional shape (12 sides), but

a much thinner baseplate (1.25" vs. 3" for the Texas HMIP) and a smaller base shaft diameter

(28.5 " vs. 32.625 ") than the HMIP considered in this study.

Page 83: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

68

Table 5.4. Texas HMIP natural frequencies vs. Iowa HMIP natural frequencies

TEXAS (FFT)

[Hz]

IOWA (FFT)

[Hz]

Difference

[%]

I MODE 0.28 0.31 10

II MODE 1.22 1.34 10

III MODE 3.20 3.33 4

Figure 5.11 shows the strain FFT (Y-axis on the right) together with the cumulative RMS (Y-axis

on the left). The RMS has a monotonic positive trend. The graph starts from zero with a steep

slope, which represents the static push of the wind, and for each natural frequency has a step that

represents how much each mode contributes in terms of strain. In the plot it can be seen that for

the record processed, the first mode contributed the most with roughly 10 microstrain cycles,

while the second mode had less than 5 microstrain cycles and occurred with a really small strain

range.

Fig. 5.11. Fourier Spectrum (FFT) and cumulative RMS of a strain gage record

Page 84: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

69

Fig. 5.12. Cumulative RMS of stress records at different triggering wind speeds

Figure 5.12 shows the cumulative RMS of stress records vs. frequency at different wind speeds.

From the plot it can be seen that the static wind push generates more stress than the natural

vibratory modes at the high wind speeds. Also it can be noticed that, among the natural modes,

the first mode contributes the most to the excitation of the pole in terms of stress. The third and

fourth modes do not contribute significantly. The second mode stress range is comparable to the

first mode induced stress range when the wind speed is close to the critical vortex shedding

speed.

The analysis results discussed in Section 5.1.1 can be used to modify the stresses measured at the

strain gage locations to the estimate the stresses at the base plate connection. Considering that

both the first and second modes contribute to the fatigue damage of the pole, an increment for the

measured stress to compute the nominal stress at the base should be between 4.5% and 11%.

Page 85: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

70

Theoretically, the increment should be differentiated depending on the level of stress considered.

For low stress cycles, which mostly happen in the 6 to 9 mph wind speed range where the second

mode vortex shedding has some effect, an increment of 11% should be applied. While for higher

wind speeds that have been shown to excite the first natural mode, the increment should be 4.5%.

It was decided to simply apply an average increment of 7% to all the stress levels which makes

the nominal stress at the connection ( ):

(Eq.5.3)

Where is the stress measured by the strain gages.

5.2.3 Wind Speed to Stress Correlation

This section focuses on the relationship between wind speed and the stress measured on the poles.

Figure 5.13 shows data points for multiple events captured on all the poles, together with

approximate expressions of the static wind push as suggested by AASHTO (Section 2.3.4), and a

second-order polynomial regression of the data points. For each data point, the x-axis coordinate

is the maximum wind speed (averaged over 3 seconds) measured during a 5 min event, and the y-

axis coordinate is the maximum measured stress from the four strain gages. No distinction was

made to see whether the stress was measured transversally or parallel to the wind direction in this

plot.

The calculations for the equivalent wind push with the AASHTO equations were carried out

considering a simplified shape of the HMIP (Fig.5.14). The lighting ring was modeled as a

rectangle with a width equal to the ring diameter (63") and a height of 11", in such a way to cover

the same surface area as the lighting apparatus. The AASHTO procedure for the calculations of

the equivalent static push was presented in Section 2.3.4.

Page 86: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

71

Fig. 5.13. Max measured Stress vs. 3 second wind speed

Page 87: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

72

Fig. 5.14. Simplified model for the equivalent wind push

AASHTO suggests the height and exposure factors Kz to account for the variation of the wind

speed along the height from the ground. The value of the exposure factor is equal to 1 at a height

of 33 ft, which is the typical height at which weather stations are located, and thus wind speed

data are available. The exposure factor has a minimum of 0.87 at the ground level and goes up to

1.37 at 150 ft from the ground.

For the gust effect factor the minimum value of 1.14 was used, while for the drag coefficient two

options were chosen:

- Cd, as shown in Fig.2.9, specified by AASHTO for a dodecagonal section (solid line)

- Cd = 1.5, which is the average value measured in the wind tunnel testing performed in the

Iowa study (MTC,2007), as shown in Figure 2.11 (doubled dash line)

Page 88: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

73

The pole cross wind surface was then divided in sections along the length and for each section the

equivalent wind pressure Pi was computed. The bending moment at the base of the pole is then

computed as:

(Eq.5.4)

Where is the equivalent wind pressure at the section – i

is the area of the section - i

is the height of the centroid of the section - i.

In Figure 5.12, it can be seen that the AASHTO equivalent push expression with a drag

coefficient is higher than the same expression with the drag coefficient suggested by AASHTO. It

can also be seen that the recorded data fit the trend of the two curves well and stayed below the

AASHTO prediction.

It can also be seen that for events with wind speed below 10 mph the stress in the structure is

higher than the static push predictions. This is because the vortex shedding effect locks in the

structure second mode of vibration in that range of wind speeds. This seems to be consistent with

what has been observed in the Iowa study, although relatively minor vortex shedding induced

stresses can be observed.

A plot of the maximum measured stress range and the corresponding average wind speed over a 1

minute period is shown in Figure 5.15. The plot shows a comparison with stresses measured at

the Iowa HMIP (Connor, Hodgson 2006) (Warpinski, 2006). It can be seen that from a wind

speed of approximately 7mph the induced stresses on the Iowa HMIPs are as high as 2 ksi, while

the stresses measured in this study are below 1 ksi. It seems then that this HMIP is less subject to

vortex shedding excitation. This is mainly due to a higher taper of the shaft and a larger cross

sectional diameter.

Page 89: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

74

Fig. 5.15. Max measured Stress vs. 1 minute wind speed

Generally it can be stated that the trend of the data points follows a monotonic second order

polynomial curve, which means that wind push from buffeting is the prevalent mechanism for the

pole.

5.3 Fatigue Assessment with Rainflow Data

This section focuses on the procedure for the fatigue life estimate using the field stress data. The

rainflow counting of the stresses provide an indication of the load paths that each pole

experienced during the monitoring period. Assuming that each pole will experience the same load

path in the future, an estimate of the fatigue life of the poles can be made using the cumulative

damage theory as expressed by Miner (1945). The following subsection provides an overview of

the cumulative damage theory.

Page 90: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

75

5.3.1 Cumulative Damage Theory

Miner’s rule states that the damage resulting from a certain stress range is proportional to the

number of cycles that fall in that stress range. The rule can be expressed by the following

equation:

(Eq.5.5)

Where is the number of cycles at the stress range i

is the number of cycles that would cause failure at the stress range i.

The rule does not consider sequence effects and the average stress of the cycle, both of which

have been seen to affect the fatigue life (NSBA, 1998). The fatigue life estimates obtained using

this method have a high variability, but they do constitute a good measure of the fatigue

performance for the poles, and thus of the hazard of a potential collapse for fatigue.

Miner also stated that a stress path composed by loading cycles of different range can be

transformed into a simple load path of constant amplitude. This equivalent stress amplitude

can be calculated as:

(Eq.5.6)

Where is the stress range of the bin i in the rainflow counting

is the ratio of the number of cycles of the bin i and the total number of cycles

.

The number of cycles to failure is calculated as:

(Eq.5.7)

Where is the equivalent stress amplitude as defined in Eq. 5.1

A is a constant that depends on the fatigue performance of the detail under consideration.

Page 91: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

76

AASHTO defines different 7 different fatigue categories as listed in Table 5.5. The best fatigue

category is A and the worst performance is category E'. For each category a value of A is

specified as indicated in Table 5.5.

Table 5.5. AASHTO A values for the fatigue performance

AASHTO FATIGUE CATEGORY A B B' C C' D E E'

A [x 108 ksi

3] 250.0 120.0 61.0 44.0 44.0 22.0 11.0 3.9

Once the value of A is defined, the estimate can then be computed as:

(Eq.5.8)

Where is the number of cycles to failure as previously defined

is the equivalent number of cycles measured during the instrumentation time

is the time length of the data set available.

It is clear that the longer the period of instrumentation, and thus the data set, the more realistic the

prediction will be. Wind activity can indeed vary depending on the season, being generally higher

in the spring time.

5.3.2 Definition of the fatigue performance

Defining the correct fatigue coefficient A for the HMIPs is of key importance in the process of

assessing the expected fatigue life. The detail used in the HMIP under consideration is a full

penetration groove welded between the tube and the transverse plate connection, welded from

both sides (without backing bar). This detail is categorized as E by the AASHTO specifications.

Category E therefore provides a good representation of the resistance for an un-cracked pole.

However, micro-cracks due to galvanizing have been found to reduce the resistance (Pool, 2010).

It is reasonable to assume the connection is an AASHTO category E′, the lowest category.

Page 92: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

77

A more accurate estimate of the resistance comes from using the cyclic load test results. A library

of six tests performed on 33" (84 cm) diameter 12-sided poles is available for comparison (Pool,

2010; Stam, 2009). Test results are available for poles without cracks, poles with initial micro-

cracks, as well as poles repaired with “field repair” and “shop repair” procedures. The plot in

Figure 5.16 shows the results of the laboratory fatigue tests in the previous research work. The

solid red circle dot represents the design that is more critical, and that is under consideration in

this study. A way to determine the A coefficient based on the performance showed in the load test

is to draw a line on the plot that passes through the data point (Figure 5.14). The expression of the

A coefficient then becomes:

(Eq.5.9)

Two estimates were made for the HMIP. The first estimate assumed AASHTO category E', and

the second estimate assumed the same capacity of the pole tested in the laboratory. Although not

presented in this thesis, two additional specimens will be fatigue tested in this research study.

One of the specimens is the base section of the pole on IH10 in El Paso, which will be taken out

of service and shipped to Ferguson Lab. Multiple ultrasonic inspections performed on this pole

within the last two years have found that the pole has cracks in 11 out of 12 of the shaft bends and

that the cracks are growing with time. This test will give the best estimate of fatigue resistance for

other poles in service that present this type of cracking.

Page 93: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

78

Fig. 5.16. Definition of the Fatigue Performance of a cracked HMIP based on load test results

5.3.2 Estimates of the fatigue life

From an observation of rainflow counting data (Fig.5.17) it can be seen that the pole on US54 is

the one with the most number of cycles for relevant level of stress. It is also the one that had the

most reliable instrumentation system in terms of continuity, and had the longest period of recoded

data. In the wintertime when the sun exposure was the shortest, the system showed a good

recharging behavior.

Most of the cycles for all the poles happen in the low stress range, smaller than 2 ksi. This means

that low stress cycles have a strong impact in the fatigue damage computation. The process of

counting stress cycles for fatigue damage evaluation uses a process referred to as rainflow

counting. In the procedure, Small stress cycle increments are defined so that cycles within a

certain range can be counted in a “bin”. For example, if a “bin” size of 0.25 ksi is selected, stress

ranges would be counted in 0-0.25 ksi, 0.25-0.5 ksi, 0.5-0.75 ksi, and so on. To understand how

Page 94: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

79

much every bin contributes to the damage of the structure a plot of versus can be

done (Fig. 5.18). It can be seen that for all the locations instrumented, the peak damage happens

at stresses below 1 ksi. The plot decreases quickly past 1 ksi for the El Paso IH-10 and Corpus

Christi HMIPs, while for the Lubbock and El Paso US-54 HMIPs cycles up to 3 ksi still possess a

significant contribution. As mentioned above, the US-54 instrumentation has been operating the

longest, ensuring a continuity of operation even in the winter months. The rainflow counting from

El Paso US54 represents the most reliable measure of the stress path exciting the structure on an

average basis.

Fig. 5.17. Rainflow Counting per day of record.

The rainflow counting was programmed to discard strain cycles smaller than 2 microstrains,

which corresponds to a stress level of 0.006 ksi. These really small cycles can indeed be

considered noise rather than actual strain measurements. AASHTO specifies a value of constant

Page 95: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

80

amplitude limit stress (ΔF) for each fatigue category. This value is a lower bound threshold for

the stress, below which cycles are not considered to cause cumulative fatigue damage in the

structure. For this specific case, since the performance of the HMIP with initial cracking is

relatively low, below every category specified by AASHTO, and because the structure is subject

to low stress cycles, it is not easy to determine a fixed value for ΔF. Therefore, rather than use a

fixed limit, increasing values of ΔF have been employed, discarding bins lower than a specified

amount from the fatigue life calculations. The following tables show the calculations of the

expected fatigue life for the poles instrumented at varying constant amplitude limit stresses. The

life estimate named “Cracked” (right end column) is computed using the load test performance,

with a value of . The plot in Figure 5.19 shows how the fatigue life estimate

(“cracked”) for each HMIP changes depending on the constant amplitude limit stress used. A

general rule to choose the appropriate limit is to take the value for which the curve changes slope

and starts growing rapidly. From the curves it is not easy to detect such a change. The El Paso

US-54 HMIP is the only pole that shows a sudden change in slope, happening between 0.4 and

0.5 ksi. It was then decided to use a constant amplitude limit stress equal to 0.5 ksi. The rows

corresponding to that stress limit have been highlighted in the tables (0.44 ksi). In many field

studies stress ranges below 0.5 ksi are often used as a limit below which stress cycles are

considered noise.

Page 96: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

81

Fig. 5.18 Fatigue damage contribution of each stress range

Page 97: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

82

Table 5.6. Fatigue life estimates for the El Paso US-54 HMIP

Table 5.7. Fatigue life estimates for the El Paso IH-10 HMIP

Constant Amplitude Limit Stress (Σ ϒi Sri3)1/3 Neq Nfail Life (E') Life (Cracked)

[ksi] [ksi] [yr] [yr]

0.06 0.48 26912148 3927155184 94 34

0.15 0.67 9971072 1466940218 95 34

0.30 0.86 4587909 692701796 97 35

0.44 1.04 2466757 391186904 102 36

0.59 1.29 1175145 203469967 111 40

0.74 1.62 538548 103169929 123 44

0.89 1.82 355735 72175559 131 47

1.03 2.01 251379 54009991 138 49

1.18 2.18 184064 42001140 147 52

1.33 2.36 137214 33411133 157 56

1.48 2.53 103273 26977139 168 60

Constant Amplitude Limit Stress (Σ ϒi Sri3)1/3 Neq Nfail Life (E') Life (Cracked)

[ksi] [ksi] [yr] [yr]

0.06 0.22 26912148 39413277120 461 165

0.15 0.42 9971072 5802153301 520 186

0.30 0.67 4587909 1457954886 610 218

0.44 0.86 2466757 683363484 700 250

0.59 1.06 1175145 370106495 825 295

0.74 1.25 538548 222819185 982 351

0.89 1.45 355735 142891048 1176 420

1.03 1.65 251379 97385906 1401 501

1.18 1.84 184064 69611570 1661 593

1.33 2.03 137214 51992336 1952 697

1.48 2.22 103273 40155822 2283 815

Page 98: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

83

Table 5.8. Fatigue life estimates for the Lubbock HMIP

Table 5.9. Fatigue life estimates for the Corpus Christi HMIP

Constant Amplitude Limit Stress (Σ ϒi Sri3)1/3 Neq Nfail Life (E') Life (Cracked)

[ksi] [ksi] [yr] [yr]

0.06 0.39 26912148 7203595397 275 98

0.15 0.72 9971072 1184167453 297 106

0.30 0.93 4587909 545654065 331 118

0.44 1.15 2466757 290050655 376 134

0.59 1.35 1175145 176943586 426 152

0.74 1.54 538548 119930302 480 171

0.89 1.71 355735 87415503 539 193

1.03 1.87 251379 66929661 609 217

1.18 2.02 184064 52771435 692 247

1.33 2.17 137214 42531759 795 284

1.48 2.33 103273 34687446 926 331

Constant Amplitude Limit Stress (Σ ϒi Sri3)1/3 Neq Nfail Life (E') Life (Cracked)

[ksi] [ksi] [yr] [yr]

0.06 0.39 26912148 7.34E+09 295 105

0.15 0.53 9971072 2.98E+09 299 107

0.30 0.69 4587909 1.32E+09 317 113

0.44 0.85 2466757 7.21E+08 352 126

0.59 1.00 1175145 4.41E+08 410 146

0.74 1.15 538548 2.88E+08 499 178

0.89 1.30 355735 1.98E+08 631 225

1.03 1.46 251379 1.41E+08 831 297

1.18 1.62 184064 1.03E+08 1122 401

1.33 1.77 137214 78411439 1534 548

1.48 1.93 103273 61034730 2148 767

Page 99: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

84

Fig. 5.19. Fatigue Life Estimate (cracked) as function of the Constant Amplitude Limit Stress for

the instrumented HMIPs

Page 100: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

85

5.4 Modification to the estimate due to short term measurement

This section suggests a modification to the fatigue life estimate to account for the fact that

measurements have not been taken throughout a whole year. The main issues of discontinuity in

the measurements were due to a low recharging power provided by the solar panels during the

winter months. The powering performance of the systems is discussed in the next section.

The following plots represent for each location instrumented the number of days of measured data

for each month (right axis), together with the average monthly wind speed at that location (left

axis). The El Paso US-54 instrumentation setup was installed in May 2010, together with the El

Paso IH-10 instrumentation, which was removed in September 2010 to be transferred to Lubbock.

Both the instrumentation setups in El Paso showed a good performance in terms of continuity and

charging power during the period of operation. The Corpus Christi and the Lubbock

instrumentations, on the other hand, showed a poor performance during the wintertime.

It is useful to observe what the average wind speed was during the days of operation as opposed

to the other months of the year. Indeed the fatigue life estimate is carried out with the assumption

that the stress history measured during the instrumentation period represents the same level of

stress that the structure will experience in the future. For this assumption to be acceptable, the

average wind activity during the monitoring period has to be close to the average yearly wind

activity.

Page 101: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

86

Fig. 5.20. Days of Operation and Average Monthly wind speed for the Corpus Christi HMIP

Fig. 5.21. Days of Operation and Average Monthly wind speed for the Lubbock HMIP

Page 102: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

87

Fig. 5.22. Days of Operation and Average Monthly wind speed for the El Paso US-54 HMIP

Fig. 5.23. Days of Operation and Average Monthly wind speed for the El Paso IH-10 HMIP

Page 103: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

88

The modification to the fatigue estimate was carried out considering the average wind speed

during the instrumentation period for each pole (Vinst) and the average wind speed throughout the

year (Vyear). Qualitatively, if the average wind speed during the instrumentation is lower than the

average wind during the whole year, it is assumed that the measurements underestimate the

excitation effect on the HMIPs. Recalling the data plot shown in Figure 5.12, the relationship

between the stress and the wind speed has been seen to be following a monotonic trend, and that

the vortex shedding effect does not have a strong impact on the low speeds. For this reason a

stress correction factor is calculated using the AASHTO equivalent static push expression (solid

line in Figure 5.12). The stress correction factor is calculated as the ratio of the two stresses

obtained inputting Vyear and Vinst in the AASHTO equation. It seems reasonable to assume that this

stresses could represent the average stress cycles that the HMIPs are subjected to, under an

average wind.

The next step is to modify the fatigue life prediction. Recalling Miner’s rule, the expression of

the number of cycles to failure, is a function of the stress range to the power of three. The

correction for the fatigue life is then obtained by elevating the stress correction factor to the

power of three. The fatigue life reduction factors for each location calculated with these

assumptions and listed in Table 5.10. It can be seen that the fatigue life estimate for the Corpus

Christi location should be reduced by 70%. This big reduction is due to the fact that the rainflow

counting available for that location where measured when the wind activity was low. The fatigue

estimate for Lubbock should be reduced by only 5%, because the average wind speed during the

instrumentation period is close to the yearly average wind speed.

Table 5.10. Modification of the fatigue life estimate for short term monitoring period

Vyear

3 [mph]

Vinst [mph]

Stress Correction Factor

Fatigue Life Reduction Factor

CORPUS CHRISTI 12.0 11.0 1.19 1.70

EL PASO - US54 8.8 8.3 1.12 1.39

EL PASO - IH10 8.8 8.3 1.11 1.36

LUBBOCK 12.4 12.3 1.02 1.05

3 Source www.weatherunderground.com

Page 104: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

89

Based on these coefficients, the fatigue life estimates previously calculated are modified (Table

5.11). These estimates could represent the expected fatigue life for all the HMIPs with the critical

design (12 sided section, 150ft tall, 80 mph design wind speed, and without external collar) that

are in service in the same locations. The fatigue life calculated assuming AASHTO category E'

would be appropriate for HMIPs with a limited amount of cracking, while the estimate using the

load test resulting capacity (cracked) would be appropriate with poles having extended cracking.

From these numbers it can be seen that the El Paso US54 pole has the highest potential for fatigue

failure. All the other locations reported values that are higher than 50 years, which is the typical

design life for these poles, thus they are less likely to encounter a fatigue failure. These estimates

represent a good measure for the fatigue failure hazard of this type of HMIPs. It can be said that

HMIPs located in the El Paso area, presenting cracking at the base connection should be

considered for a replacement or a repair. One of the reasons why the El Paso area could be more

critical for the fatigue performance of these poles is that the average windspeed throughout the

year has a peak of occurrence in the 6 to 8 mph range (Fig. 4.3). In this location the II mode

vortex shedding could have a bigger impact on the excitation of the structure, causing the HMIPs

in that area to experience more cycles than the ones in other locations.

Table 5.11. Modification of the fatigue life estimate for short term monitoring period

Location Fatigue Life (E')

[years] Fatigue Life (Cracked)

[years]

EL PASO - US54 73 26

EL PASO - IH10 515 184

LUBBOCK 358 128

CORPUS CHRISTI 207 74

Page 105: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

90

Chapter 6 Conclusions

6.1 Introduction

This thesis documented a research investigation on the fatigue performance of high mast

illumination poles. The work presented is part of a larger study that has included laboratory tests,

monitoring the behavior of HMIP sections during galvanizing, computational studies, and

monitoring of HMIPs in the field. The primary focus of this thesis was the field monitoring of

existing HMIPs in the state of Texas and a small portion of the computational studies.

Measurements of stress and wind speed on high mast illumination poles where used to determine

the dynamic behavior under wind loading and to estimate their remaining fatigue life. Chapter 2

of the thesis provided a summary of background information from previous studies on the

subject. In Chapter 3 a description of the instrumentation used for the field monitoring was

provided. Chapter 4 showed preliminary data regarding the HMIP locations. In Chapter 5 the

observation of the measured data, a study on the dynamic behavior of the poles, the fatigue life

estimate, as well as the supporting finite element models were discussed. This final chapter

presents the conclusions for this research work.

6.2 Finite Element Analyses

The computational studies presented in this thesis consisted of results from a modal analysis from

a finite element model of the poles subjected to wind and gravity loading. Two models were used

consisting of a line element model and a shell element model. The support conditions of the

poles consisted of restraints from translation and rotation at the base of the HMIP.

- The modal analysis performed on the frame cantilever model of the HMIP provided good

estimates for the natural periods of the poles. The comparison of the computed natural

periods with the Fourier spectrum of the measured stresses showed that the error was less

than 10% on the third natural mode.

- The second order bending moment at the base connection generated by differential

heating on the HMIP was computed to be 43k-in.

Page 106: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

91

- The analysis performed on the shell finite element model showed that the support

conditions can affect the stress concentration at the shaft to base plate connection of the

HMIP, in such a way that the fatigue life can reduce by a factor of 1.8.

6.3 Field Monitoring Data

- Vortex shedding takes place at a wind speed close to 7 mph, exciting the second natural

mode. The maximum nominal stress range at the pole connection caused by vortex

shedding is less than 1 ksi. When compared to the HMIP designs in other states, the

TXDOT pole geometry has a higher taper of the shaft, which mitigates the effect of

vortex shedding on the structure.

- The El Paso pole on US-54 was found to have a remaining fatigue life lower than the

design life, which is 50 years. Observing the recording from weather stations in El Paso it

was observed that the average daily wind speed is close to 7 mph.

- The fatigue life estimate showed that the equivalent stress amplitude for all the HMIP

locations was close to 1ksi.

- Both vortex shedding and buffeting contribute to the fatigue damage of the poles.

6.4 Future Work

- The fatigue testing that will be performed on the HMIP taken out of service from the El

Paso, IH-10 location will provide an indication of the fatigue performance of a pole that

is heavily cracked in the field. This information will allow another estimates of the

fatigue life for the poles instrumented.

- Considering that the equivalent stress measured on the poles was close to 1 ksi, the new

load testing should be performed with a lower stress amplitude than what was used in the

previous phases. The minimum stress amplitude that could be used will be controlled by

the time required to reach the expected number of cycles to failure at the maximum

loading frequency that the hydraulic actuators can deliver.

- From the inspection of the cracks after the fatigue test of this pole it will be possible to

determine if the cracks were growing in the field as it seems to be from the multiple

Page 107: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

92

Ultrasonic inspections performed. The cracks will be cut open to determine whether their

nature is fatigue or thermal, from the galvanizing process.

6.5 Recommendations to HMIP Owners

- An important aspect of proper HMIP installation is that the top and bottom anchor nuts

on the baseplate are properly tightened and tack welded, because that has been found to

affect the fatigue life of the poles.

- Based on the results of the fatigue life estimates it is recommended that a replacement or

a repair of the poles that are found to be heavily cracked should be considered by the

owners.

Page 108: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

93

Appendix A: Texas Department of Transportation’s Standard Plans for High Mast

Illumination Poles

Page 109: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

94

Page 110: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

95

Page 111: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

96

References

1. American Association of State Highway and Transportation Officials, AASHTO Standard

Specifications for Structural Supports for Highway Signs, Luminaries and Traffic Signals,

Fourth Edition, AASHTO, Washington, D.C., 2001.

2. American Society of Civil Engineers, Minimum Design Loads for Buildings and Other

Structures. ASCE7-10, 2010.

3. Anderson, T.H., Fatigue Life Investigation of Traffic Mast-Arm Connection Details, M.S.

Thesis, Department of Civil Engineering, The University of Texas at Austin, August 2007.

4. Chang, B., Phares, B.M., Sarkar, P.P., and Wipf, T.J., Development of a Procedure for

Fatigue Design of Slender Support Structures Subjected to Wind-Induced Vibration,

Transportation Research Record: Journal of the Transportation Research Board, No. 2131,

Transportation Research Board of the National Academies, Washington, D.C., 2009, pp. 23–

33. DOI: 10.3141/2131-03

5. Clough, R.W., Penzien, J., Dynamics of Structures, Third Edition, Computers and Structures

Inc., Berkeley, CA, 2003.

6. Connor, R.J., Hodgson, I.C., Field Instrumentation, Testing, and Long-Term Monitoring of

High-Mast Lighting Towers in the State of Iowa, Pennsylvania Department of Transportation

Final Report, November 2006

7. Downing, S.D., Socie, D.F., Simple Rainflow Counting Algorithms, International Journal of

Fatigue, Volume 4, Issue 1, pp. 31-40, 1982.

8. Fisher, J.W., Kulak, G.L., and Smith, I.F.C., A Fatigue Primer for Structural Engineers,

NSBA/AISC, Chicago, IL, 1998.

Page 112: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

97

9. High Mast Illumination Poles. Standard Plans, Texas Department of Transportation

(TXDOT), August 1995.

10. Kinstler, T.J., Current Knowledge of the Cracking of Steels During Galvanizing.

GalvaScience LLC, Springfield, Alabama, 2006

11. Koenigs, et al., Fatigue Resistance of Traffic Signal Mast-Arm Connections. Texas

Department of Transportation, Research Report 4178-2, Center for Transportation Research,

August 2003.

12. Lynn, P., Fuerst, W., Introductory digital signal processing with computer applications,

Second Edition, John Wiley, 1998.

13. Miner, M.A., Cumulative Damage in Fatigue, Journal of Applied Mechanics, pp. 159-164,

1945.

14. Pool, C.S., Effect of Galvanization on the Fatigue Strength of High Mast Illumination Poles,

The University of Texas at Austin, Master’s Thesis, 2010.

15. Rios, C.A., Fatigue Performance of Multi-Sided High-Mast Lighting Towers. M.S. Thesis,

Department of Civil Engineering, The University of Texas at Austin, May 2007.

16. Stam, A.P., Fatigue Performance of Base Plate Connections Used in High Mast Lighting

Towers. M.S. Thesis, Department of Civil Engineering, The University of Texas at Austin,

May 2009

17. Warpinski, M.K., The Effect of Base Connection Geometry on the Fatigue Performance of

Welded Socket Connections in Multi-sided High-mast Lighting Towers, Department of Civil

Engineering, Leigh University, May 2006

Page 113: Copyright by Luca Magenes 2011fsel.engr.utexas.edu › pdfs › MAGENES-THESIS.pdf · Luca Magenes, M.S.E. The University of Texas at Austin, 2011 Supervisor: Todd A. Helwig Failures

98

Vita

Luca Magenes was born in Pavia, Italy. After completing his Bachelor’s in Civil Engineering at

The University of Pavia in 2008, he enrolled in the University of Texas at Austin to pursue a

Master’s Degree.

Contact: [email protected]

This thesis was typed by the author.